time-dependent mechanics of living cells

166
London Centre for Nanotechnology Department of Mechanical Engineering Time-Dependent Mechanics of Living Cells Emadaldin Moeendarbary Dissertation submitted for the degree of Doctor of Philosophy University College London September 2012

Upload: others

Post on 11-Sep-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Time-Dependent Mechanics of Living Cells

London Centre for Nanotechnology

Department of Mechanical Engineering

Time-Dependent Mechanics of

Living Cells

Emadaldin Moeendarbary

Dissertation submitted for the degree of

Doctor of Philosophy

University College London

September 2012

Page 2: Time-Dependent Mechanics of Living Cells

Declaration

I, Emadaldin Moeendarbary confirm that the work presented in this thesis is my own.

Where information has been derived from other sources, I confirm that this has been

indicated in the thesis.

Page 3: Time-Dependent Mechanics of Living Cells

.یسپاس گزارم که به من نیروی اندیشیدن داد ، پروردگارا

:به تقدیم

یاری کرده اند ، مرا در دانش آموختنواره م پدر و مادرم که ه

و

.همسرم ژانت که همراه و پشتیبان من در به پایان رسانیدن این کار بود

Page 4: Time-Dependent Mechanics of Living Cells

IV

Acknowledgements

I am sincerely grateful to my main supervisor Dr. Guillaume Charras and second advisor

Dr. Eleanor Stride for their ‎kind guidance and continuous support over the past three

years. I would particularly like to thank Dr. Charras for offering me this project,

welcoming me to carry out this research in his lab and introducing me to biological

techniques. I would also like to extend my grateful appreciation to Prof. L. Mahadevan

for providing great insights on poroelasticity and his valuable inputs to this work.

I appreciate very much the helps of Dr. Marco Fritzsche, Dr. Andrew Harris, Dr. Carl

Leung, and Dr. Richard Thorogate from London Centre for Nanotechnology. I gratefully

acknowledge support of Prof. Nicos Ladommatos and Dr. William Suen from Department

of Mechanical Engineering. I am thankful to our collaborators Prof. Adrian Thrasher and

Dr. Dale Moulding from Institute of Child Health.

I would like to acknowledge the financial support from the Engineering and Physical

Sciences Research Council (EPSRC) through the prestigious Dorothy Hodgkins

Postgraduate Award.

Thank you my friends, Marco, Alireza, Andrew, Cesar and Miia. You are always so

helpful and encouraging.

Thank you, Judy and Guy Whitmarsh for your kind supports.

And many many thanks to my teachers and professors from elementary, junior and high

school to ‎university, who encouraged me to continuously learn and provided me with the

foundations for ‎my academic career.‎

Page 5: Time-Dependent Mechanics of Living Cells

V

Abstract

Cells sense and generate both internal and external forces. They resist and transmit these

forces to the cell interior or to other cells. Moreover a variety of cellular responses are

excited and influenced by transducing mechanical stimulations into chemical signals that

lead to changes in cellular behaviour. The cytoplasm represents the largest part of the cell

by volume and hence its rheology sets the maximum rate at which any cellular shape

change can occur.

To date, the cytoplasm has generally been modelled as a single-phase viscoelastic

material; however, recent experimental evidence suggests that its rheology can be

described more effectively using a poroelastic formulation in which the cytoplasm is

considered to be a biphasic system constituted of a porous elastic solid meshwork

(cytoskeleton, organelles, macromolecules) bathing in an interstitial fluid (cytosol). In

this framework, a single parameter, the poroelastic diffusion constant pD , sets cellular

rheology scaling as 2~ /pD Ex m with E the elastic modulus, x the hydraulic pore size,

and m the cytosolic viscosity. Though this poroelastic view of the cell is a conceptually

attractive model, direct supporting evidence has been lacking. In this work, such evidence

is presented and the concept of a poroelastic cell is validated to explain cellular rheology

at physiologically relevant time-scales.

In this work, the functional form of stress relaxation in response to rapid application of a

localised force by atomic force microscopy microindentation is examined in detail and it

is shown that at short time-scales cellular relaxations are poroelastic. Then, pD is

measured in cells by fitting experimental stress relaxation curves to the theoretical model.

Page 6: Time-Dependent Mechanics of Living Cells

VI

Next, using indentation tests in conjunction with osmotic perturbations, the validity of the

predicted scaling of pD with pore size is qualitatively verified. Using chemical and

genetic perturbations, it is shown that cytoplasmic rheology depends strongly on the

integrity of the actin cytoskeleton but not on microtubules or intermediate filaments.

Finally, comparison of scaling of viscoelastic and poroelastic models suggests that short-

time scale viscoelasticity might be due to water redistribution within the cytoplasm and a

simple scaling relating cytoplasmic viscosity to cellular microstructure is provided.

Page 7: Time-Dependent Mechanics of Living Cells

7

Contents

DECLARATION ............................................................................................................................ II

ACKNOWLEDGEMENTS ..........................................................................................................IV

ABSTRACT .................................................................................................................................... V

CONTENTS .................................................................................................................................... 7

LIST OF FIGURES ...................................................................................................................... 12

LIST OF TABLES ........................................................................................................................ 15

SUPPLEMENTARY VIDEO ...................................................................................................... 15

CHAPTER 1 INTRODUCTION ........................................................................................... 16

1.1 A Microscopic view: biological structure of cell ................................................................ 16

1.2 The role of mechanics in cell function .............................................................................. 24

1.3 Experimental techniques for measuring cellular mechanical properties .......................... 27

1.3.1 Atomic force microscopy ............................................................................................. 29

1.4 Cell mechanics and rheology ............................................................................................ 32

1.4.1 The origin of cellular mechanical and rheological properties ...................................... 32

1.4.2 Cell as a continuum media ........................................................................................... 32

1.4.3 Universality in cell mechanics ...................................................................................... 34

Page 8: Time-Dependent Mechanics of Living Cells

8

1.4.4 Top-down approaches to studying cell rheology ......................................................... 36

1.4.5 A bottom up approach: The cell as a dynamic network of polymers ........................... 39

1.4.6 Biphasic models of cells: a coarse-grained bottom up approach ................................ 43

1.5 Aims and motivation ........................................................................................................ 44

1.6 Outline of the following chapters..................................................................................... 45

CHAPTER 2 THEORETICAL OVERVIEW OF CONTINUUM MECHANICAL

THEORIES FOR LIVING CELLS .............................................................................................. 47

2.1 Theory of linear isotropic elasticity .................................................................................. 47

2.1.1 Hertzian contact mechanics: Indentation of elastic material ...................................... 48

2.2 Linear viscoelasticity ........................................................................................................ 51

2.2.1 Storage and loss moduli ............................................................................................... 52

2.2.2 Standard linear solid model ......................................................................................... 53

2.2.3 Power law models ........................................................................................................ 54

2.3 Poroelasticity ................................................................................................................... 56

2.3.1 Theory of linear isotropic poroelasticity ...................................................................... 56

2.3.2 Diffusion equation ........................................................................................................ 57

2.3.3 Scaling law between poroelastic properties and microstructural parameters ............ 58

2.3.4 Poroelastic swelling/shrinking of gels under osmotic stress ........................................ 59

2.4 Basic 1D solutions for fundamental poroelasticity problems ........................................... 61

2.4.1 Uniaxial strain (confined compression) ........................................................................ 61

2.4.2 One dimensional consolidation .................................................................................... 62

2.5 Indentation of a poroelastic material ............................................................................... 67

2.5.1 Indentation of a thin layer of poroelastic material ...................................................... 69

CHAPTER 3 MATERIALS AND METHODS .................................................................... 71

3.1 Cell culture, generation of cell lines, transduction, and molecular biology ....................... 71

Page 9: Time-Dependent Mechanics of Living Cells

9

3.2 Fluorescence and confocal imaging .................................................................................. 72

3.3 Cell volume measurements .............................................................................................. 73

3.4 Disrupting the cytoskeleton ............................................................................................. 75

3.4.1 Pharmacological treatments ........................................................................................ 75

3.4.2 Genetic treatments ...................................................................................................... 75

3.5 Visualising cytoplasmic F-actin ......................................................................................... 76

3.6 Atomic force microscopy indentation experiments .......................................................... 78

3.6.1 Measuring indentation depth and cellular elastic modulus......................................... 79

3.6.2 Measurement of the poroelastic diffusion coefficient ................................................ 81

3.6.3 Determining the apparent cellular viscoelasticity ........................................................ 82

3.6.4 Measuring cell height ................................................................................................... 83

3.6.5 Spatial probing of cells ................................................................................................. 84

3.7 Microinjection and imaging of quantum dots .................................................................. 85

3.8 Rescaling of force relaxation curves ................................................................................. 86

3.9 Polyacrylamide hydrogels ................................................................................................ 87

3.10 Particle tracking using defocused images of fluorescent beads ................................... 88

3.10.1 Experimental protocol and calibration setup .......................................................... 88

3.10.2 Radial projection ..................................................................................................... 90

3.11 FRAP experiments ....................................................................................................... 91

3.11.1 Effective bleach radius ............................................................................................ 93

3.11.2 FRAP analysis ........................................................................................................... 93

3.11.3 Estimation of cortical F-actin turnover half-time .................................................... 94

3.12 Data processing, curve fitting and statistical analysis .................................................. 94

CHAPTER 4 THE CYTOPLASM OF LIVING CELLS BEHAVES AS A POROELASTIC

MATERIAL 96

Page 10: Time-Dependent Mechanics of Living Cells

10

1.4 Indetation test for characterisation of poroelastic material ........................................ 97

4.1.1 Establishing the experimental conditions to probe poroelasticity .............................. 97

4.1.2 Poroelasticity is the dominant mechanism of force-relaxation in hydrogels ............... 99

4.1.3 Cellular force-relaxation at short timescales is poroelastic ....................................... 102

4.2 Kinetics of whole cell swelling/shrinking........................................................................ 106

4.2.1 Swelling/shrinkage of a constrained poroelastic material ......................................... 108

4.2.2 Poroelasticity can explain the kinetics of cellular swelling/shrinking ........................ 111

4.3 Discussion and conclusions ............................................................................................ 111

CHAPTER 5 POROELASTIC PROPERTIES OF CYTOPLASM: OSMOTIC

PERTURBATIONS AND THE ROLE OF CYTOSKELETON ‎ .......................................... 120

5.1 Determination of cellular elastic, viscoelastic and poroelastic parameters from AFM

measurements ........................................................................................................................ 121

5.2 Poroelasticity can predict changes in stress relaxation in response to volume changes . 123

5.3 Changes in cell volume result in changes in cytoplasmic pore size. ................................ 127

5.3.1 Hyperosmotic shock halts movement of quantum dots inside cytoplasm ................ 128

5.3.2 Translational diffusion of fluorescent probes affected by osmotic perturbations .... 129

5.4 Estimation of the hydraulic pore size from experimental measurements of Dp. ............. 133

5.5 Spatial variations in poroelastic properties .................................................................... 133

5.6 Poroelastic properties are influenced by cytoskeletal integrity ...................................... 135

5.6.1 Chemical treatments .................................................................................................. 135

5.6.2 Genetic modification .................................................................................................. 136

5.7 Discussion and conclusions ............................................................................................ 141

CHAPTER 6 GENERAL DISCUSSION AND FUTURE WORK .................................. 148

Page 11: Time-Dependent Mechanics of Living Cells

11

6.1 Further discussion .......................................................................................................... 148

6.1.1 Cell rheology and its complexity ............................................................................... 148

6.1.2 Why poroelasticity? ................................................................................................... 149

6.1.3 Direct physiological relevance .................................................................................. 152

6.2 Future work ................................................................................................................... 153

6.2.1 Stress and pressure wave propagation in cells......................................................... 153

6.2.2 Cells as multi-layered poroelastic materials ............................................................. 155

6.3 Conclusions .................................................................................................................... 157

REFERENCES........................................................................................................................... 158

Page 12: Time-Dependent Mechanics of Living Cells

12

List of Figures

Figure ‎1-1 Living cells ...................................................................................................... 17

Figure ‎1-2 The three major types of filaments that constitute the cytoskeleton. .............. 20

Figure ‎1-3 Generation of force in the actin-rich cortex moves a cell. .............................. 22

Figure ‎1-4 Binding proteins induce structural phase transitions to polymerised solution of

purified filaments. ............................................................................................................. 23

Figure ‎1-5 Cells sense and respond to application of forces. ............................................ 25

Figure ‎1-6 Common experimental techniques for measurement of cell rheology. ........... 28

Figure ‎1-7 Atomic force microscopy imaging of the cell surface..................................... 31

Figure ‎1-8 Universal phenomenological behaviours of the cell and models of cell

rheology. ........................................................................................................................... 34

Figure ‎1-9 Schematic representation of four different models of cell rheology. .............. 35

Figure ‎1-10 Dynamic network of filaments determine the cell rheology. ........................ 41

Figure ‎2-1 The geometry of contact of non-conforming bodies. ...................................... 49

Figure ‎2-2 The standard linear solid viscoelastic model. .................................................. 55

Figure ‎2-3 Compression of a porous solid with an interstitial fluid by a porous plunger. 62

Figure ‎2-4 Creep response of a poroelastic half-space. .................................................... 64

Figure ‎2-5 Creep response of a finite domain. .................................................................. 65

Figure ‎2-6 Stress relaxation response of a poroelastic half-space. ................................... 67

Figure ‎2-7 Spherical indentation and force relaxation a poroelastic material of finite

thickness. ........................................................................................................................... 68

Figure ‎3-1 Cell volume measurement. .............................................................................. 74

Figure ‎3-2 AFM experimental setup. ................................................................................ 77

Figure ‎3-3 AFM force-distance curve. .............................................................................. 80

Figure ‎3-4 Calculating the height of cell at contact point. ................................................ 83

Page 13: Time-Dependent Mechanics of Living Cells

13

Figure ‎3-5 Spatial AFM measurement .............................................................................. 85

Figure ‎3-6 High resolution defocusing microscopy of fluorescent bead. ......................... 89

Figure ‎3-7 Estimation of the effective bleach radius in HeLa cells FRAP experiments... 92

Figure ‎4-1 Stress relaxation of HeLa cells in response to AFM microindentation ........... 98

Figure ‎4-2 Force-relaxation curves acquired with different indentation depths on

hydrogel. ......................................................................................................................... 100

Figure ‎4-3 Normalisation of force-relaxation curves acquired on hydrogel. .................. 101

Figure ‎4-4 Force-relaxation curves acquired with different indentation depths on cells. 103

Figure ‎4-5 Normalisation of force-relaxation curves acquired on MDCK cells. ............ 104

Figure ‎4-6 Normalisation of force-relaxation curves acquired on HeLa cells. ............... 105

Figure ‎4-7 Experimental setup to investigate the poroelastic behaviour of cells during

swelling/shrinkage. ......................................................................................................... 107

Figure ‎4-8 Kinetics of cell swelling/shrinking in HeLa cells ......................................... 110

Figure ‎4-9 Schematic representation of intracellular water movements in different

experimental setups. ........................................................................................................ 118

Figure ‎5-1 Fitting force-relaxation curves. ..................................................................... 122

Figure ‎5-2 Poroelastic and elastic properties change in response to changes in cell

volume. ........................................................................................................................... 125

Figure ‎5-3 Effects of hypoosmotic or hyperosmotic shock on HeLa cells F-actin structural

organization..................................................................................................................... 127

Figure ‎5-4 Fluorescence recovery after photobleaching in HeLa cells. .......................... 129

Figure ‎5-5 Changes in cell volume change cytoplasmic pore size. ................................ 132

Figure ‎5-6 Spatial distribution of cellular rheological properties. .................................. 134

Figure ‎5-7 Cellular distribution of cytoskeletal fibres. ................................................... 136

Figure ‎5-8 Ectopic polymerization of F-actin due to CA-WASp. .................................. 137

Figure ‎5-9 Effect of drug treatments and genetic perturbations on poroelastic properties.

........................................................................................................................................ 138

Figure ‎5-10 Cell rheology and disease. ........................................................................... 140

Figure ‎5-11 Effect of volume changes, chemical and genetic treatments on the lumped

pore size. ......................................................................................................................... 143

Page 14: Time-Dependent Mechanics of Living Cells

14

Figure ‎5-12 Schematic representation of the cytoplasm and effects of different cellular

perturbations. .................................................................................................................. 147

Figure ‎6-1 Microstructural parameters and mesoscopic length scale set the poroelastic

time scale. ....................................................................................................................... 151

Figure ‎6-2 Stress and pressure wave propagation in cells .............................................. 154

Page 15: Time-Dependent Mechanics of Living Cells

15

List of tables

Table ‎5-1 Effects of osmotic perturbations on translational diffusion coefficients. ....... 131

Supplementary video

Effects of osmotic shock on the movement of microinjected PEG-passivated quantum

dots in HeLa cell.

Movement of PEG-passivated quantum dots microinjected in a HeLa cell in isoosmotic

conditions (left) and in hyperosmotic conditions (right). Both movies are 120 frames long

totalling 18 s and are single confocal sections. In isoosmotic conditions, quantum dots

move freely (left); whereas in hyperosmotic conditions, quantum dots are immobile

because they are trapped within the cytoplasmic mesh (right). Scale bars = 5 μm.

Page 16: Time-Dependent Mechanics of Living Cells

16

Chapter 1

Introduction

Equation Chapter (Next) Section 1

1.1 A Microscopic view: biological structure of cell

Living cells, the elementary units of life, all carry out similar basic functions and they

display a common architecture despite their various shapes, sizes and internal

complexities (Figure ‎1-1). In addition to basic functions such as proliferation, protein

synthesis, molecule transport and energy conversion, cells may also perform more

specialised roles that pertain to the organs that they are part of. Prokaryotic cells, such as

bacteria, lack membrane-bounded nuclei while eukaryotic cells contain membrane-

bounded compartments (including nuclei and other organelles) that perform specific

metabolic activities and energy conversion. Plant and animal cells are eukaryotes and are

typically larger (~ 5 to 100 µm) and more complex than prokaryotic cells (~ 1 to 10 µm).

The internal and external environments of the eukaryotic cells are separated by a plasma

membrane which is a selectively-permeable barrier consisting of a lipid bilayer with

Page 17: Time-Dependent Mechanics of Living Cells

17

embedded proteins. The work described in this thesis is focused on the animal cell whose

fluid-like plasma membrane cannot alone adopt complex cellular morphologies [1, 2]).

The material inside the cell membrane, excluding the nucleus, is generally referred to as

the cytoplasm which consists of a liquid phase, the cytosol (water, salts, small proteins),

and a solid phase (cytoskeleton, organelles, ribosomes, etc). A general diagram of animal

cell organisation is given in Figure ‎1-1B.

Figure ‎1-1 Living cells

(A) Coloured scanning electron micrograph (SEM) of a HeLa cancer cell undergoing cell division. Source:

http://www.sciencephoto.com/media/137830/enlarge. (B) Schematic showing the generic cytoplasm

organisation and nucleus of an animal cell. The main cellular structures and organelles are listed in the right

panel. Source: http://en.wikipedia.org/wiki/File:Biological_cell.svg.

Page 18: Time-Dependent Mechanics of Living Cells

18

The cellular skeleton (cytoskeleton) is a complex network of three major types of

polymeric fibre of varying size and rigidity extending throughout the cytoplasm. It

consists of cable-like actin filaments with diameters of ~ 7 nm, rope-like intermediate

filaments with diameters of ~ 10 nm and pipe-like microtubules with diameters of ~ 20

nm [1-3] (see Figure ‎1-2). The cytoskeleton is responsible for maintenance of cell shape,

cell migration, resistance to externally applied mechanical forces, and cell division.

Furthermore it controls and facilitates the transport of intracellular particles and

organization of cell contents.

Actin filaments are known to be the most important polymer network for the mechanics

of the cytoskeleton. They are organized in a variety of structures such as networks of

filaments within the cytoplasm, the cortex under the plasma membrane, the lamellipodia

at the cell edges, the filopodia and contractile stress fibres. The subunit of an actin

filament (F-actin) is a globular actin protein (G-actin) which accounts for about 5-10

percent of the total cellular protein content [4, 5]. A typical actin filament has a

longitudinal elasticity of ~ 2 GPa and a persistence length of ~ 15 µm which is larger than

its typical contour length of ~ 2 μm. Actin filaments have a distinct polarity originating

from the structural asymmetry of G-actin and filaments exhibit a faster polymerisation

rate at one end (plus barbed-end) compared to the other end (minus pointed-end).

αβ tubulin heterodimers organize into long hollow cylinders to form microtubules which

are the second major constituent of the cytoskeleton [4]. Similar to F-actin, microtubules

are polar filaments because of the structural arrangement of their subunits [2].

Microtubules are straight and rigid with turnovers on the order of minutes [6], and they

play a key role in chromosomal distribution during cell division. They typically radiate

randomly from the centrosome (which is a microtubule organizing centre (MTOC) near

the nucleus) towards the periphery of the cell with their (–) ends proximal to the MTOC

and their (+) ends directed away from the MTOC. The tubular structure of microtubules

gives rise to a higher bending stiffness than actin filaments despite their similar effective

longitudinal Young’s moduli [2]. The persistence length of microtubules is ~ 5 mm which

Page 19: Time-Dependent Mechanics of Living Cells

19

is ~ 1000 fold larger than their typical length ~ 5 µm and the typical length of a cell: ~

50 ‎µm‎. Microtubules are known to be the most significant compressive-load bearing

elements of the cytoskeleton [7].

As the most abundant cytoskeletal filaments, intermediate filaments are located both

inside the cytoplasm and the nucleus and form a large and more diverse family of highly

a -helical proteins. Unlike actin filaments and microtubules subunit proteins, these a ‎-

helical ‎proteins integrate into more diverse sequences with greatly varying molecular

weights to form different types of intermediate filaments such as: keratins that are

abundant in epithelial cells, lamins that are involved in the formation of a structure called

the lamina which is located underneath nuclear membrane and helps stabilise the nuclear

envelope, and vimentin filaments that are expressed mainly in mesenchymal cells [3].

Intermediate filaments, typically have a diameter of ~ 10 nm which is in between the

diameter of actin filaments (7 nm) and microtubules (25 nm).They have a persistence

length of ~ 1 µm that is smaller /comparable to their length giving them more flexibility.

Intermediate filaments extend out from around the nucleus throughout the cytoplasm and

it is thought that they are mechanically incorporated with other parts of cytoskeleton

through specialised crosslinking proteins to reinforce the integrity of the cell [8].

Intermediate filaments are more stable than actin filaments and microtubules (for instance

keratin filaments have a turnover of ~ 30 min [9]) and play a major role in cell-cell

attachments in epithelial tissue [10].

Page 20: Time-Dependent Mechanics of Living Cells

20

Figure ‎1-2 The three major types of filaments that constitute the cytoskeleton.

(A,B,C -I): The electron micrographs and lattice structure of filaments: source:[3] (A,B,C –II): Schematic of

spatial distribution of cytoskeleton filaments. (A,B,C –III): Distribution of cytoskeletal constituents in live

HeLa cells. All images are maximum projections of confocal stacks. Green indicates the localisation of the

GFP-tagged construct. The nucleus is shown in blue in all images. (A-III) F-actin was enriched in cell

protrusions but was also present in the cell body. (B-III) GFP-α-tubulin was homogenously localised in the

cell body. A-III & B-III were acquired by L. Valon. ‎(C-III) Intermediate filaments (keratin 18) were

homogenously distributed in the cell body. Scale bars = 10 μm.

Page 21: Time-Dependent Mechanics of Living Cells

21

As briefly described in the previous paragraphs, each cytoskeletal filament is constructed

from specific protein subunits. There exists a continual flux of constituent subunits

onto/from both ends of filament which results in polymerisation/depolymerisation. In the

steady state, polymerisation and depolymerisation are balanced and as a result the

filament maintains a given length, a phenomenon known as treadmilling (see

Figure ‎1-3C-III for actin filament). Under normal conditions, the rate of this

polymerisation/depolymerisation (turnover) is different for each of the three major types

of cytoskeletal filament. For instance, intermediate filaments have a much slower rate of

turnover (~ tens of minutes) than actin filaments (~ tens of seconds). Many active features

in the cell such as force generation are in part due to these energy consuming turnover

processes. For instance, actin polymerization [11-13] constructs the leading edge of the

cell, a two dimensional sheet-like structure called the lamellipodium, that drives cell

protrusion (see Figure ‎1-1 and Figure ‎1-3).

The above three major filament types are the key parts of the cytoskeleton. To build the

cytoskeletal network architecture and define the mechanics of the cell, these primary

filaments interact with one another, cross-link and bind with themselves and connect to

other parts of cytoplasm (plasma/nuclear membrane, organelles, etc) through various

types of linking proteins and molecular motors. Motor proteins convert the energy

released by ATP hydrolysis into mechanical work ‎to dynamically interact with the

cytoskeleton and to facilitate many different cellular functions such as force generation

and locomotion, cell division and intracellular transport. For instance during cell

movement, myosin II motor proteins move along cytoskeletal filaments and generate

contractile forces necessary for contraction of the cell rear (see Figure ‎1-3D and

Figure ‎1-4). Crosslinking proteins organize cytoskeletal filaments into entangled and

bundled complex scaffolds. Hence, they help determine the network architecture leading

to different cellular mechanical functions [14]. For example, the crosslinking protein α-

actinin is one of the most abundant acting binding proteins and it crosslinks actin

filaments into bundles to provide contractility in cytoskeletal assemblies such as stress

Page 22: Time-Dependent Mechanics of Living Cells

22

fibres [15] and the contractile ring [16] (Figure ‎1-4). Other crosslinkers, such as filamin,

organise actin filaments into a crosslinked gel-like network (Figure ‎1-4).

Figure ‎1-3 Generation of force in the actin-rich cortex moves a cell.

(A) Scanning electron microscope image of a fish keratocyte. (B) Phase-contrast image of the keratocyte

moving downward with a speed of ~ 15 µm.s-1. (C-I) Distribution of cytoskeletal filaments with actin

fillaments in brown, microtubules in green and intermediate filaments in blue. (C-II) Electron micrograph of

the F-actin network in the lamellipodium. (C-III) Treadmilling: polymerisation/depolymerisation of actin

filaments by reversible addition of actin subunits to the free ends of the filament. One end of the filament

elongates (plus end) while the other end shrinks (minus end). At steady state, addition and loss are balanced.

(D) Schematic diagram of how cells move forward. Actin polymerisation at the leading edge of the cell

extends a protrusion in the direction of movement. The leading edge adheres to the substrate in front and

adhesions at the rear of the cell are detached. The contractile forces generated by the acto-myosin at the cell

rear push the whole cell body forward. Source:[3].

Page 23: Time-Dependent Mechanics of Living Cells

23

Figure ‎1-4 Binding proteins induce structural phase transitions to polymerised solution of purified filaments.

(A) Addition of binding proteins with different properties such as type, organisation and concentration results

in the formation networks with different microstructural organisations. Adapted from [17]. (B) Effects of

crosslinkers and myosin II thick filaments on the network topology of actin filaments. (B-I) Entangled

network of purified actin filaments. (B-II) Addition of α-actinin crosslinker proteins changes the shape and

architecture of the actin filament ‎network. (C) Addition of myosin II ,motor proteins leads to further

contraction and rearrangement ‎of the crosslinked F-actin network‎. Source: [18].

Page 24: Time-Dependent Mechanics of Living Cells

24

1.2 The role of mechanics in cell function

Living cells sense their environment and generate internal and external forces [19]. They

resist and transmit these forces to the cell interior or other cells through chemical and

physical signals. Complex sensory machineries located on different cellular sites such as

primary cilia, stretch-‎gated ion channels and focal adhesions are responsible for sensing

mechanical ‎stimulations ‎(arising from application of different types of forces such ‎as

tensional, compressional, pressure and ‎shear forces) generated internally or applied

externally on ‎different parts of the cell‎ and eventually transducing these cues into

biochemical signals (see Figure ‎1-5A) ‎[20, 21]. More fascinatingly, in a closed-loop

feedback manner these mechanosensed biochemical signals trigger energy-consuming

chemical processes relying on ATP hydrolysis that cells employ to modify their own

behaviour and generate forces to respond to and resist the initial mechanical cues.

Mechanotransductory processes greatly influence and control a variety of cellular

responses that lead to changes in cellular behaviour and function such as changes in cell

morphology, lineage and fate [22, 23]. For example, in the musculoskeletal system, actin

and myosin molecular motors produce active contraction in muscle cells in response to

external electrical stimuli such as nerve influxes. Inner hair cells transduce the fluid

oscillation generated in the cochlea by sound waves into electrical signals that are

transmitted to the brain. Vascular endothelial cells bear hemodynamic forces, such as

blood flow shear stress and pressure, and convert these applied stresses into intracellular

signals that influence cellular functions such as proliferation, migration, permeability, and

gene expression [24]. More interestingly, fluid shear stresses can modulate endothelial

structure and function (see Figure ‎1-5B). Indeed endothelial cells, through rearrangement

of their cytoskeletal structures and active remodelling processes, change their shape and

align themselves with the direction of the flow to control the magnitude of shear forces

they are exposed to [24]. One outstanding challenge in cell biology is to understand how

Page 25: Time-Dependent Mechanics of Living Cells

25

mechanical cues can be transduced into biochemical cues and how biochemical changes

in the cell can give rise to changes in cellular mechanics or the forces applied by cells.

Figure ‎1-5 Cells sense and respond to application of forces.

(A-I) Application of different types of forces on cell such as shear forces induced by fluid flow over the cell,

tensile forces arising from the extracellular matrix (ECM) and internal compressional or tensional forces

induced and borne by the cellular cytoskeleton. (A-II) Diagram of a mechanosensor at the focal adhesions. At

the focal adhesion site, the internal cytoskeletal forces are balanced by external forces distributed on the ECM

(blue). On the external side of the cell membrane, ECM is attached to focal adhesions (multicolored array

of ‎proteins) through integrins (brown). On the internal side of the cell, membrane actin microfilaments (red)

are anchored into focal adhesions. (III) The internal and external forces are passed through the

mechanosensory site where forces can be sensed, controlled and regulated in a closed-loop feedback. Source:

[20] (B) Endothelial cell morphology transformed by longterm exposure to fluid shear stress. Increasing the

level of shear stress from 0 to > 1.5 Pa causes alignment of bovine aortic endothelial cells (shown under static

culture conditions in I) in the direction of fluid flow after 24 hours (shown in II).Source: [25].

Entanglement (physical) and crosslinking (biochemical) of filamentous polymer networks

inside ‎cells maintain cell shape and govern cellular mechanical properties such as

elasticity. ‎Molecular motors and crosslinkers contract the cytoskeletal polymer network

Page 26: Time-Dependent Mechanics of Living Cells

26

and induce ‎pre-stresses. Active mechanical forces generated from chemical reactions

(such as active ‎polymerisation of filaments or molecular motors) enforce morphogenetic

changes such as cell ‎rounding, cytokinesis, cell spreading, or cell movement. During

these gross morphogenetic ‎changes the cytoskeleton provides the force for

morphogenesis but the maximal rate at which ‎shape change can occur is dictated by the

rate at which the whole cell can be deformed. ‎Hence, ‎the dynamic mechanical properties

(rheology) of cytoplasm, which are poorly understood and are ‎the main topic of this

work, play a major role in setting the rate at which any morphogenetic ‎event can occur.

Furthermore, cells detect, react, and adapt to external mechanical ‎stresses by regulation of

microscopic processes that changes their mechanics. However, in the absence of an in

depth understanding of cell rheology, the transduction of external stresses into

intracellular mechanical changes is poorly understood, making the identification of

the ‎physical parameters that are detected biochemically largely speculative. ‎

Although the correlation between the mechanical properties of tissues and diseases

has ‎been recognized quite widely for a long time, the ‎connection between the mechanics

of living cells (at single cell level) and disease is poorly understood. Indeed, the cell

phenotype in health or disease can be affected by the rheology of the cell. One clinical

example suggests that changes in cell rheology can have consequences for the health of

patients: some patients with a low neutrophil count exhibit a constitutively active

mutation in the Wiskott Aldrich syndrome protein (CA-WASp) that results in over-

activation of actin polymerisation through dysregulated activation of the Arp2/3 complex,

leading to delays in several stages of mitosis [26, 27]. This overabundance of cytoplasmic

F-actin increases cellular apparent viscosity resulting in delays in all phases of mitosis

and causing kinetic defects in mitosis (Moulding et al, Blood, 2012). Further to this

example, there are other reports that identified different mechanical properties for

cancerous cells compared to those of normal cells [28-31] which illustrate the importance

of studying cell rheology. The ultimate goal is to combine theoretical, experimental and

Page 27: Time-Dependent Mechanics of Living Cells

27

computational approaches to construct models for a realistic description of the various

types of cellular mechanical behaviours based on their microscopic cytoskeletal

arrangement, something necessary to enable a true physical understanding of the stress

fields present and generated in tissues and developing embryos during physiological

functions.

1.3 Experimental techniques for measuring cellular mechanical

properties

Depending on the length scale of the cellular structure under investigation and the

required spatial and temporal resolution, many experimental techniques have been

utilized to measure the mechanical properties of the different cellular components. In

general, most methods measure the static/dynamic response of the cellular structure to the

application of controlled forces or deformations. Other techniques track the motion of

endogenous or embedded particles of various sizes to study the rheology of the cell. The

important point is that due to the heterogeneity and complexity of the cell and the

cytoskeleton, the submicrometre-scale measurement techniques can lead to considerably

different evaluations of mechanical properties compared to bulk (several micrometre

scale) measurement techniques. One of the main challenges is to find a universal

framework under which the measured macroscopic properties can be interpreted to obtain

realistic information about the dynamics of the microstructure of the cell or vice versa,

i.e. to estimate the bulk rheological properties from the measured microscopic properties

[32].

Page 28: Time-Dependent Mechanics of Living Cells

28

Figure ‎1-6 Common experimental techniques for measurement of cell rheology.

Several experimental methods such as magnetic twisting cytometry [33, 34], magnetic

tweezers [35], optical tweezers [36], substrate cell stretchers [37, 38], shear flow

rheometry [39, 40] (see Figure ‎1-6) and atomic force microscopy (AFM) (see the

following paragraph and Figure ‎1-7) have been utilized to perturb small regions of the

cell or deform an entire cell to investigate cell rheology. An alternative technique is the

use of passive methods, such as traction force microscopy that involves quantification of

cellular traction forces using different detection mechanisms such as micropillar arrays

[41] or substrate-embedded beads [42]. One more recently applied ‎‎technique is particle-

Page 29: Time-Dependent Mechanics of Living Cells

29

tracking microrheology that tracks motion of embedded or attached tracer particles

present within the system to extract cellular mechanical properties [43]. Depending on the

cell type, the experimental condition and the probing technique diverse ranges of cellular

rheological properties have been reported and various mechanical models ‎(some of which

are described in the following sections)‎ were applied ‎to interpret quantitative

measurements of cell mechanical properties. Further to the need for significant

improvement in mechanical models, the experimental methods also ‎required to be

improved to measure local/global mechanical properties of single cells ‎more accurately.

Some of these important features include: the ability to impose/measure ‎deformations and

forces on micro and nano scales, perform dynamic experiments, integrate with ‎other

experimental tools such as confocal microscopy and maintain cell viability for a

sufficient ‎period of time‎ [4, 44-46].

1.3.1 Atomic force microscopy

In my studies, I utilized the Atomic force microscope (AFM), a very high resolution form

of scanning probe microscope that was invented in 1986 [47] and has emerged rapidly as

a versatile tool to study biological samples over the past two decades‎ [48]. This allowed

me to acquire topographic images of cells (Figure ‎1-7D), probe rigidity [49] and

characterisation of viscoelastic properties of different cell lines (Moulding et al, Blood,

2012) and extracellular gel matrices (Calvo et al, Nature Cell Biology, under review).

Topographic AFM measurements involve a tip connected to a micro-fabricated cantilever

being scanned over the surface of a sample in a series of horizontal sweeps (see

Figure ‎1-7A). A laser beam from a solid state diode is reflected off the back of the

cantilever and collected by photodiodes. Perturbations in the movement of the tip due to

surface topography result in changes in bending of the cantilever and consequently of the

reflection path of the laser beam which is sensed via photodiodes. A piezo-electric

ceramic in a feedback loop is used to move the cantilever up and down to maintain a

constant bending of the cantilever. In contact mode, the tip is touching the surface while

Page 30: Time-Dependent Mechanics of Living Cells

30

sweeping across it. As the deflection of the cantilever is kept constant, the surface

topography is reconstructed in 3D from the piezo movement and the planar trajectory of

the cantilever (see Figure ‎1-7A). The cantilever vertical and lateral deflections provide

information about the sample height changes and can map the surface distribution of

different chemical functionalities [50].

In addition to high resolution imaging of biological structures, AFM has been used

extensively to assess the mechanical properties of biological materials such as elasticity

and viscoelasticity [51-54]. As will be explained later in Chapter 3.7, AFM indentation

was employed in this thesis to characterise the rheological properties of cytoplasm.

Furthermore AFM can be integrated with other techniques such as defocusing microscopy

to investigate stress propagation inside cells and this idea will be described in the last

chapter.

Page 31: Time-Dependent Mechanics of Living Cells

31

Figure ‎1-7 Atomic force microscopy imaging of the cell surface.

(A) Schematic diagram of an AFM cantilever tip scanning a surface. Vertical cantilever motion is controlled

by a piezoelectric ceramic through a feedback loop. Bending of the cantilever alters the laser beam reflection

path which is sensed by photodiodes. (B-I,II) SEM images of AFM cantilevers with pyramidal (I) and glued

spherical (I) tips. In (II) a glass microbead was glued onto the cantilever. (C) Phase contrast image of an AFM

cantilever probing HeLa cells cultured on a glass coverslip. (D-I) Contact mode image of HeLa cells shown

in (C), imaged in cell culture medium with a pyramid tipped cantilever. (D-II) Three dimensional

reconstruction of cell topography. Scale bars = 10 µm.

Page 32: Time-Dependent Mechanics of Living Cells

32

1.4 Cell mechanics and rheology

1.4.1 The origin of cellular mechanical and rheological properties

As a first step to studying cellular mechanical and rheological properties, the cell can be

simply considered as a conventional engineering material [55]. Simple conventional

materials exhibit very simple frequency-dependent responses: for normal solids the shear

modulus ‎is independent of frequency and for liquids it is simply proportional to

frequency. However, materials ‎(including cells) ‎with more complex hierarchical

structures and a high degree of heterogeneity can exhibit a variety of complex responses

depending on strength, frequency and spatial application of deformations and forces.

Being such a complex material, the cell displays viscoelasticity, i.e. it behaves like an

elastic solid over short time scales with a finite shear modulus whereas it acts as a viscous

liquid at long time scales [56, 57]. In intermediate regimes, both the storage and shear

moduli govern the mechanical response of the cell. At different hierarchical levels,

individual microstructural elements of the cytoskeleton are perturbed by application of

force and deformed at microscopic scales to accommodate macroscopic deformations.

Furthermore, processes such as continuous turnover of cytoskeletal fibres,

association/dissociation of crosslinkers ‎and activity of molecular motors mean that the

cell is an active biological material [56] that exhibits astonishing rheological behaviours

by coupling active and passive biochemical and mechanical processes. Indeed, time-

dependent measurements of cellular properties reveal a spectrum of relaxation times

which are strongly influenced by the size, stability, geometry, and flexibility of cross-

linkers, active motions, and changes in filament structure due to turnover [45]. ‎

1.4.2 Cell as a continuum media

Experiments on cells have shown that the cytoskeleton is the main determinant of cellular

mechanical properties. Hence, much effort has concentrated on describing cytoskeletal

rheology. As will be described in the following, wide ranges of top-down and bottom-up

Page 33: Time-Dependent Mechanics of Living Cells

33

theoretical models have been proposed to describe the rheology of the cytoskeleton and

the cellular environment. A robust model is one that can incorporate the minimum

necessary information about internal structure, spatial granularity, heterogeneity and the

active features unique to cells to predict long-wavelength and long-timescale

redistribution of stress [46, 56]. However the search for such a comprehensive model is

still ongoing. Indeed, even a description of the cell as a multi-layered composite (such as

an elastic membrane tightly bound to the subcortical actin gel, a more granular gel-like

inner layer, a more fluid layer in the cell interior and a stiff nucleus) without

incorporating the internal complexity of each layer and active processes still remains a

very complex system to investigate. Indeed, experimentally measured rheological

properties are highly dependent on the size of the probe as well as its connectivity and

interaction with the cellular structures [58]. For example, during microindentation

experiments with a micron-sized bead, the induced deformations on the cell are

significantly larger than the mesh size of the cytoskeletal network such that continuum

viscoelastic models can be safely used without concern for the heterogeneous distribution

of filamentous proteins in the cytoskeleton. Therefore, operating at sufficiently large

length scales, cells can be considered as continuum media in which the contribution of the

microarchitecture can be coarse-grained through constitutive laws. Despite the limited

ability of continuum models to explain complex cytoskeletal structures, to date they are

still the most efficient means of describing experimental observations on the micrometre

scale and to model stress/strain transmission through the cell. These models are

particularly useful when associated with coarse-grained relations giving scaling between

continuum level rheology and cellular microstructure. Considering the cell as a single

phase material, several types of viscoelastic models have been employed to predict the

biomechanical behaviour of cells [44]. On the other hand, multiphasic models such as

biphasic poroelasticity have been applied to investigate the effects of the different

constitutive phases on cell rheology [4, 59].

Page 34: Time-Dependent Mechanics of Living Cells

34

Figure ‎1-8 Universal phenomenological behaviours of the cell and models of cell rheology.

(A) The frequency response of cells measured with several measurement techniques. Two master curves were

obtained after rescaling the results from different rheological measurements. Adapted from: [58]. (B)

Spontaneous retraction of a single actin stress fibre upon severing ‎with a laser nanoscissor shows ‎existence of

prestress in the cytosketal bundles. Scale bar = 2 ‎µm‎‎.Source:[60]. (C) Anomalous diffusion and response of

the cell to stretch. Spontaneous movements of beads attached firmly to the cell (C-I, II) show

intermittent ‎dynamics. Adapted from [61]. (C-III) The mean square displacement (MSD) of a bead anchored

to the ‎cytoskeleton exhibits anomalous diffusion dynamics (subdiffusive at short time intervals

and ‎superdiffusive at longer time intervals). ‎Red curve ‎indicates the MSD of the bead for a non-stretched cell

and other curves show the MSD of the bead in response to a global stretch measured at different waiting

times ( wt ) after stretch cessation. Scale bars = 10 µm.‎‎Source: [62].‎

1.4.3 Universality in cell mechanics

Cell mechanical studies over the years have revealed a rich phenomenological ‎landscape

of rheological behaviours that are dependent upon probe geometry, loading ‎protocol and

loading frequency [58, 63, 64]. More recent mechanical measurements on eukaryotic

cells agree on the presence of four universal cellular phenomenological behaviours [65]:

1) Cell rheology is scale free: plots of the frequency response of many cell types on log-

log scale display the same shape and follow a weak power law spanning several decades

Page 35: Time-Dependent Mechanics of Living Cells

35

of frequency [58]. ‎2) Cells are prestressed: mechanical stresses generated continuously by

the internal activity of ‎actomyosin or applied externally on the cell are

counterbalanced ‎by the tensional/compressional state of the cytoskeleton‎‎ [66]. ‎‎3) The

fluctuation dissipation theorem (FDT) breaks down and diffusion is anomalous [61]: The

spontaneous motion of endogenous particles or embedded/attached beads present within

the system ‎does not follow the Stokes-Einstein relationship (see section 1.4.4.2 for

explanation of this relationship). ‎‎4) Stiffness and dissipation are altered by stretch:

Application of stretch significantly perturbs the rheological properties of the cell and

depending on the experimental condition the cell can exhibit different behaviours such as

stress stiffening, fluidisation and rejuvenation [62]. See Figure ‎1-8 for brief illustration of

these universal behaviours.

Figure ‎1-9 Schematic representation of four different models of cell rheology.

(A) Examples of linear viscoelastic models. (B) Tensegrity [66] (C) Soft glassy rheology [67]. (D) Network

of semiflexible filaments models (with or without effects of crosslinkers and molecular motors).

Page 36: Time-Dependent Mechanics of Living Cells

36

Several top-down (linear viscoelasticity, tensegrity, soft glassy rheology) and bottom-

up ‎(networks of semiflexible polymers) ‎theoretical models (see Figure ‎1-9) have been

successfully applied to explain the observed phenomenological universal behaviours [64].

However there is still no unifying theory to explain the physical mechanisms that govern

these observed universal cellular behaviours. As will be described briefly in the following

sections, linear viscoelasticity, immobilized colloids/ soft glasses and the network of

semiflexible polymers are the most widely used models that have been proposed to study

cell mechanics.

1.4.4 Top-down approaches to studying cell rheology

1.4.4.1 Linear viscoelastic models

Elastic and viscoelastic constitutive laws have been applied to study the response of the

whole cell to external forces and to investigate the stress and strain distribution inside the

cytoplasm. Whereas the static mechanical properties of cells (elasticity) have been

studied in depth, the time-dependent mechanical properties of cells have received

significantly less attention. The majority of the work to date utilises an empirical

viscoelastic description of cells that assumes the cytoplasm is a single phase homogenous

material [32, 53, 68, 69]. The fundamental properties of simple elastic and viscoelastic

materials are explained briefly in Chapter 2. The most commonly used viscoelastic

models consider the whole cell or part of it as a homogeneous continuum, where the

smallest length scale is significantly larger than the dimensions of the microstructural

constituents. Examples of such models are cortical shell-liquid core, elastic/viscoelastic

solid, and power-law models [44, 64]. The cortical shell-liquid core and

elastic/viscoelastic solid models can generate good fits to experimental data by

introducing a finite number of elastic and viscous elements ‎(springs and

dashpots) ‎coupled in series or parallel leading to exponential decay functions with a finite

number of relaxation times [44]. One example of such a model used to fit stress

Page 37: Time-Dependent Mechanics of Living Cells

37

relaxation is illustrated in Chapter 2.2.2. In contrast to spring-dashpot models, power law

models do not have any characteristic relaxation time and cannot be easily described

using mechanical analogs. Power law structural damping models have been applied more

widely in recent microrheological experiments that measure the viscoelastic response of

the cells over a broader range of frequencies and times. Despite their widespread use, the

major limitation of these models is that they are not mechanistic, fail to relate the

measured rheological properties to structural or biological parameters within the cell, and

thus cannot predict changes in rheology due to microstructural changes.

1.4.4.2 Cell as a soft glassy material: macromolecular crowding and anomalous

diffusion

A large body of more recent research has found that some of the cellular rheological

behaviours are empirically similar to the rheology of soft materials such as foams,

emulsions, pastes, and slurries. Following some experimental observations, it was

proposed that cells could be considered as soft glassy materials [67]. Indeed, measuring

the fluctuations of particles within the cytoplasm revealed that in many cases they exhibit

a larger random amplitude of fluctuations than expected and a greater degree of

directionality than could be developed solely from thermal fluctuations [70]. In a normal

diffusion process (Brownian motion), the mean squared displacement (MSD) of a particle

with size a evolves linearly in time t , 2 6 Tr D t , where TD is the particle

translational diffusion coefficient ( TD can be calculated using the Stokes-Einstein

relationship /6T BD k T apm where m is the viscosity of the medium). However, a

semi-empirical equation 2 ~r ta can be used to describe particle fluctuations when it

exhibits non-Brownian behaviour as is frequently observed in cellular environments [71,

72]. This deviation from the Stokes-Einstein relationship has been attributed to reaction

forces from active cytosketal processes [73] and crowding of a large number of

macromolecules inside the cytoplasm such as mobile intracellular globular proteins and

Page 38: Time-Dependent Mechanics of Living Cells

38

other fixed obstacles like cytoskeletal filaments and organelles [74, 75] that reduce the

available solvent volume and provide barriers to particle Brownian motion [76-78].

Based on these observations, it was proposed that the high concentration of different

proteins in the cytoplasm can lead to liquid crystal and colloidal behaviours that can be

interpreted in terms of the ‘‘soft glassy rheology’’(SGR) model [67]. Crowded colloidal

suspensions or soft glasses exhibit a weak power-law rheology corresponding to a

continuous spectrum of relaxation times. Similar to this, the dynamic modulus of cells

scales with frequency as a weak power law valid over a wide spectrum of time [33, 79].

This semi-empirical SGR behaviour is in contrast to the rheology of semiflexible network

of filaments that generally predict a frequency independent dynamic modulus at low

frequencies [58].

As a conceptual model, SGR explains how macroscopic rheological responses are linked

to localized structural rearrangements originating from structural disorder and

metastability. Briefly, the SGR system consists of particles that are trapped in energy

landscapes arising from their interactions with surrounding neighbours if there exists

sufficient crowding. In such a system, thermal energy is not sufficient to drive structural

rearrangement and as a consequence out of equilibrium trapping occurs. With time

remodelling/rearrangement (micro-reconfiguration) happens when particles escape the

energy barriers of their neighbours and jump from one metastable state to another

reaching a more stable state with relaxation rates slower than any exponential process

[61]. In such a system, injecting agitational energy sourced from non-thermal origins

(such as mechanical shear or ATP-dependent conformational changes of proteins [62])

liberates particles from the energy cages in which they are trapped and facilitates

structural rearrangements causing the material to flow.

Page 39: Time-Dependent Mechanics of Living Cells

39

1.4.5 A bottom up approach: The cell as a dynamic network of polymers

At very short time scales (high frequencies), the cellular viscoelastic dynamic modulus

scales with frequency with a universal exponent of 0.75 [80]. This behaviour is analogous

to the observed rheology of semiflexible polymer networks [81-84]. This analogy as well

as the physical picture of the cell constituted of cytosketal filaments (Figure ‎1-10A)

suggest that the viscoelastic cellular properties originate mainly from the entropic and

enthalpic interactions between cytoskeletal structures which largely depend on the

architecture and mechanical properties of the constituent cytoskeletal filaments [85].

The flexibility of polymers can be characterised by two length scales, the persistence

length pl and the contour length L [86]. The persistence length or the length of thermal

flexibility pl is the length scale over which thermal bending fluctuations become

appreciable and can change the direction of the filament, /p f f Bl E I k T where

f fE Ik is the bending modulus of a single filament ‎(or flexural rigidity defined by

single filament elastic modulus fE and its cross-sectional second moment of inertia fI )

and Bk T the thermal energy. A filament is called rigid when pL l and flexible when

pL l (Figure ‎1-10B). However, most cytoskeletal gels are networks of semiflexible

biopolymer chains with the persistence length of individual chains comparable to their

contour length ~ pL l . In such networks, the elastic and dissipative properties originate

from entropic (where energy is stored entropically due to the reduction of the number of

accessible chain spatial configurations in the network [87]) and enthalpic (where energy

is stored in extensional and bending modes of filaments) processes that involve filaments

and also the interaction of filaments with the viscous fluid that they are bathed in [86, 88-

90]. Furthermore, crosslinking of these semiflexible filaments by specialised (rigid or

flexible) proteins at very short distances cl along the length of each individual filament

significantly perturbs the network rheology. Indeed, depending on the properties of

Page 40: Time-Dependent Mechanics of Living Cells

40

crosslinkers such as length, flexibility and concentration different nonlinear elastic effects

can be observed [64].

For purified gels of cytoskeletal filaments such as actin gels, based on a microscopic

picture of a network of entangled and crosslinked semiflexible filaments, various

mechanical models have introduced different scaling relationships relating viscoelastic

properties to microstructural parameters. The persistence length pl , geometric mesh size

l and a characteristic length chL have been shown to determine the viscoelasticity of

these purified gels. For an actin gel with an actin concentration Ac and a monomeric actin

size a , the geometrical mesh size (a measure of the nearest distance between two

neighbouring monomers forming part of different filaments) is given by ~1/ Aacl

[91]. In dense network regimes ( pa ll ) the elastic modulus scales as

2 2 3~ B p eE k T l ll where el is the entanglement length (the confining length scale that

restrict topological motion of neighbouring filaments). In this entangled network the

characteristic length ch eL l is the entanglement length el and the physical effect of

crosslinking from entanglements is to constrain the motion of filaments to a tube-like

region, with a diameter 3/2 1/2~e e pd l l , surrounded by other filaments (Figure ‎1-10C-I).

On the other hand for a densely crosslinked gel, the mesh size ( scales the same as the

entanglement length) is the characteristic length of the network ~ch eL l l

(Figure ‎1-10C-II) and the elastic modulus scales as 5 2~ B pE k T ll .

Page 41: Time-Dependent Mechanics of Living Cells

41

Figure ‎1-10 Dynamic network of filaments determine the cell rheology.

(A) Image of actin filaments stained with Alexa Fluor 647 phalloidin in a COS-7 cell taken by dual stochastic

optical reconstruction microscopy (STORM). The z-position from the substrate is colour coded with respect

to the scaling shown in the colour scale bar. Scale bars = 2 μm. Source: [92]. (B) The contour length L of a

polymer compared to its persistence length pl changes the dynamics of polymer networks from flexible to

semiflexible and rigid regimes (from left to right respectively). (C) Schematic representation of minimal

models for the dynamics of networks of polymers. Semiflexible polymer chains such as actin filaments

exhibit complex dynamics under steric entanglements or crosslinks. (C-I) Tube picture of polymer dynamics.

Two length scales, l the mesh size and el the entanglement length (or ed the tube diameter), set the network

dynamics. (C-II) For densely crosslinked polymers the mesh size l or entanglement length ~el l sets the

dynamics of network. Source: [32].

Page 42: Time-Dependent Mechanics of Living Cells

42

The experimentally determined elastic moduli of cells were found to be several orders of

magnitude larger than those measured in in vitro studies of stress-free F-actin gels and

this suggested that the main elastic properties of cells could not result solely from

semiflexible networks in a stress free state, such as networks of filamentous actin [93].

This was explained by the fact that unlike most conventional materials, the viscoelastic

response of semiflexible networks of biopolymers is highly nonlinear (tension or

deformation-dependent) and thus the prestressed state of filaments in the cytoskeleton

could result in the high measured elasticities for cells [94]. In addition to the external

application of tensional forces, active intracellular processes such as contraction mediated

by myosin motor proteins could also generate this prestress. The prestress is transmitted

through the cytoplasm by the actin cytoskeleton and balanced by adjacent cells and the

extracellular matrix. Furthermore as suggested in the tensegrity model [66, 95], the

prestress distribution inside the cytoplasm might be partly balanced by compression of

other cytoskeletal filaments such as microtubules.

Considering the cytoskeletal network as a physical gel of purified cytoskeletal filaments

is a ‎first step for studying mechanical properties of the cell but yet too simplistic.

Indeed ‎cytoskeletal filaments and proteins ‎‎continuously consume energy released from

ATP ‎hydrolysis that turns the system into a non-equilibrium state. This ATP energy is

involved in ‎filament polymerisation/depolymerisation, active association/dissociation of

crosslinkers ‎and the activity of molecular motors. New theoretical models have been

recently proposed to study ‎such non-equilibrium systems [57, 96]. One category of

models considers the detailed dynamics of ‎filaments and their interactions with molecular

motors and builds up ‎mechanical properties of ‎the filamentous system using microscopic

equations and numerical simulations [97]. Alternatively ‎hydrodynamic models describe

the behaviour of the mixture by establishing ‎phenomenological ‎relations between a few

coarse-grained variables [96, 98]. ‎

Page 43: Time-Dependent Mechanics of Living Cells

43

1.4.6 Biphasic models of cells: a coarse-grained bottom up approach

Considering a cell to be a single phase material is counterintuitive given that more than

60 percent of the cellular content is water. In mammalian cells, most studies view water

solely as a solvent and an adaptive component of the cell that engages in a wide range of

biomolecular interactions [99]. However less attention has been given to the significance

of water in the dynamics of the cytoskeleton, its role in cellular morphology, and motility.

Recent experimental work showed the presence of transient pressure gradients inside cells

and suggested that these could be explained by the biphasic nature of cytoplasm [100-

103]. As a consequence, a biphasic description of cells was proposed based on

poroelasticity (or biphasic theory), in which the cytoplasm is biphasic consisting of a

porous elastic solid meshwork (cytoskeleton, organelles, macromolecules) bathed in an

interstitial fluid (cytosol) [100, 101, 103-105]. In this framework, the viscoelastic

properties of the cell are a manifestation of the time-scale needed for redistribution of

intracellular fluids in response to applied mechanical stresses and the response of the cell

to force application depends on a single experimental parameter: the poroelastic diffusion

constant pD [106], with larger poroelastic diffusion constants corresponding to more

rapid stress relaxations. A minimal scaling law 2~ /pD Ex m relates the diffusion

constant to the drained elastic modulus of the solid matrix E , the pore size of the solid

matrix x , and the viscosity of the cytosol m [100]. Therefore, contrary to viscoelastic

models, the dynamics of cellular deformation in response to stress derived from

poroelasticity can be described using measurable cellular parameters, allowing

changes of rheology with E , x , and m to be predicted which makes this framework

particularly appealing conceptually.

Page 44: Time-Dependent Mechanics of Living Cells

44

1.5 Aims and motivation

One of the most striking features of eukaryotic cells is their capacity to change shape in

response to environmental or intrinsic cues driven in large part by their actomyosin

cytoskeleton. In studies of cellular morphogenesis, the cytoplasm is generally viewed as

an innocuous backdrop enabling diffusion of signalling proteins. This view fails to

account for the fact that the time-dependent mechanical properties (or rheology) of the

cell are determined by the cytoplasm because it forms the largest part of the cell by

volume. During gross morphogenetic changes such as cell rounding, cytokinesis, cell

spreading, or cell movement, the cytoskeleton provides the force for morphogenesis but

the maximal rate at which shape change can occur is dictated by the rate at which the

cytoplasm can be deformed. Using a combination of techniques from molecular cell

biology and nanotechnology, my aim is to investigate the biphasic nature of animal cells

and its implications for cell rheology. I examine the contribution of intracellular water

redistribution to cellular rheology at time-scales relevant to cell physiology (0.5-10s) and

investigate how the cytoskeleton and macromolecular crowding interact to set cellular

rheology.

Experimental measurements of cellular viscosity are generally interpreted with models

that either consider the cell as a single phase homogenous material or describe cell

rheology with a continuous spectrum of relaxation times. These models lack predictive

power because rheological parameters cannot be related to cellular structural and

constitutive parameters. Recent studies [100-102, 107] suggest that water plays an

important role in cellular mechanical properties. Based on the physical nature of

cytoplasm, it was suggested that cells are like fluid-filled sponges. One very interesting

implication of the fluid-sponge model is that cellular rheology depends on only one

parameter, the poroelastic diffusion constant, that scales with three structural and

constitutive parameters: the sponge elasticity, the sponge pore size, and the viscosity of

Page 45: Time-Dependent Mechanics of Living Cells

45

the interstitial fluid. Though this is a conceptually attractive model, direct experimental

evidence to validate this model has been lacking.

In this thesis, such evidence is presented and I seek to study cell rheology within the

framework of poroelasticity. I ask whether cell rheology exhibits a poroelastic behaviour

similar to well characterised poroelastic materials such as hydrogels and then I investigate

the validity of a poroelastic scaling law by perturbing cellular parameters that are

important in setting poroelastic properties. To study cell rheology two experimental

approaches were optimised: AFM micro-indentation stress-relaxation tests and whole cell

swelling/shrinking experiments. Cellular stress relaxations were measured by rapidly

applying a localized force onto cells via an AFM cantilever and monitoring the ensuing

stress relaxation over time (Figure ‎3-2). As an alternative approach, the ability of the

poroelastic theory to predict the swelling/shrinking kinetics of the cells was investigated

by applying sudden osmotic perturbations and monitoring associated dynamics by

defocusing microscopy (Figure ‎4-7).

1.6 Outline of the following chapters

In Chapter 2, the basic principles of elastic, viscoelastic and poroelastic continuum

mechanical theories are introduced. In particular in relation to indentation experiments,

Hertzian contact mechanics is briefly overviewed. The theory of linear isotropic

poroelasticity is explained more comprehensively followed by the solutions of some

fundamental problems useful in interpreting the presented experimental measurements.

In Chapter 3, the techniques and methodologies employed to determine the poroelastic

properties of cells are presented. These include cell preparation, chemical and genetic

treatments, AFM indentation experiments, particle tracking as well as fluorescent and

brightfield microscopy.

Page 46: Time-Dependent Mechanics of Living Cells

46

In Chapter 4, the ability of the poroelastic theory to describe cellular stress relaxation in

response to rapid application of a localised force by AFM are investigated. The cellular

force-relaxation curves are compared with stress relaxation curves acquired on well

characterised poroelastic hydrogels and the functional form of these curves is discussed in

detail. Also in this chapter, the kinetics of cell swelling/shrinking is studied by employing

poroelasticity to describe the dynamics of cell volume changes induced via sudden

osmotic perturbations and monitored by defocusing microscopy.

In Chapter 5, using indentation tests in conjunction with osmotic perturbations, the

validity of the predicted scaling of pD with pore size is verified qualitatively. Also using

chemical and genetic perturbations, the dependence of cytoplasmic rheology on the

cytoskeleton is explored. In this chapter, the significance and the implications of the

experimental results for understanding the biological determinants of cellular rheology

are discussed.

In Chapter 6, the conclusions from this work are summarised and the implications of

poroelasticity for cell mechanics are discussed further. Also, prospective work and further

experiments that can be performed in the future are presented, including other

experimental techniques to investigate time-dependent mechanical properties of cells.

Page 47: Time-Dependent Mechanics of Living Cells

47

Chapter 2

Theoretical overview of continuum

mechanical theories for living cells

Equation Chapter (Next) Section 1

2.1 Theory of linear isotropic elasticity

The elastic framework is the simplest continuum formulation that provides the

relationship between deformations/strains and forces/stresses. Linear elasticity is the very

simplified version of finite strain theory in which the material strains ( ε strain tensor) are

considered to be infinitesimal (small deformations) and have a linear relationship with

imposed stresses (σ stress tensor). From Newton’s second law, the equation of motion for

such material transforms into

σdiv r rB u (2.1)

where r is the density of material, B body force per unit mass, u is the vector of solid

displacement for small deformations:

Page 48: Time-Dependent Mechanics of Living Cells

48

ε .0.5 Tu u (2.2)

The div and designate the divergence and gradient operators, respectively. For an

isotropic and homogeneous material, the stress-strain relationship is expressed through

the constitutive equations:

σ ε ε 2 tr( )m l I (2.3)

where m and l are Lamé’s constants, I the identity tensor and tr the trace operator. The

first Lame constant m is equivalent to the shear modulus G . The following equations

define the relationships between Lame’s constants , the elastic Young’s modulus E and

the Poisson ratio n [108]

,2(1 )2

.1 2

Em

nmn

ln

(2.4)

Using equations (2.1), (2.2), and (2.3), the stress field can be computed from a prescribed

displacement field and boundary conditions. On the other hand, compatibility conditions

[108] are required to derive a unique displacement field knowing the components of the

stress tensor. Typically in most materials, including cells, the body forces and inertial

terms can be neglected and thus the equation (2.1) transforms into equilibrium equation,

σdiv 0.

2.1.1 Hertzian contact mechanics: Indentation of elastic material

Here the Landau and Lifshitz [109] approach is presented for studying the contact

problem for spherical linear isotropic elastic bodies which is known as Hertz’s contact

problem. It can be shown that when two smooth surfaces contact each other without

deformation (Figure ‎2-1A) their separation in the z direction h can be expressed in terms

of their principal relative radii of curvature. For the contact of two spheres with radii of

1R and 2R , the separation takes the form:

2 2

2 2*

1 2

( )1 1 1( ) ,

2 2x y

h x yR R R

(2.5)

Page 49: Time-Dependent Mechanics of Living Cells

49

where x and y are coordinates lying on the tangent plane of the contact with the contact

point as the origin and 1 2 1 2* /R RR R R the relative radius of curvature. Now let

us apply a total compressive load of F to the touching spheres such that the centres of

the two spheres approach one another by an amount 1 2d d d . The separation distance

between two bodies at a given height z , can be calculated from geometrical

considerations shown in Figure ‎2-1B:

'1 2( )h h w wd (2.6)

where 1( , )w x y and 2( , )w x y are the elastic deformations for surface points on each body

respectively. It is trivial that within the contact area ' 0h and outside the contact area

' 0h so that the surfaces do not overlap.

Figure ‎2-1 The geometry of contact of non-conforming bodies.

(A) Unloaded case where the curved objects are just touching. (B) Loaded case where the curved objects are

pressed into each other. Source: [110].

Considering the Hertzian conditions (i.e. continuous, non-conforming and frictionless

surfaces, small elastic strains where the radius of contact circle a is much smaller than

the relative radius of curvature *a R and that each sphere can be considered as an

Page 50: Time-Dependent Mechanics of Living Cells

50

elastic half-space), the normal displacement of each surface 1( , )w x y or 2( , )w x y , can be

written in terms of the normal component of pressure p on the surface:

' '

' '' 2 ' 2

( , )1( , ) .

2 ( ) ( )

p x yw x y dx dy

x x y y

npm

(2.7)

Within the contact area, ' 0h , substituting the above equation into the geometrical

relationship (2.6) yields:

' '1 2 ' '

' 2 ' 21 2

2 2

1 2

( , )1 1 12 ( ) ( )

1 1 1( ).

2

p x ydx dy

x x y y

x yR R

n np m m

d (2.8)

Solving (2.8) by applying potential theory [109], results in the following relationships for

the pressure field distribution and the contact radius

23

( ) 1 ,2 m

rp r p

a (2.9)

1 2 1 2

1 2 1 2

3 1 18 m

RRa p

R Rp n n

m m (2.10)

where 2 2r x y and mp is the average pressure over the contact area

2

0( )2 .

a

mF p r rdr p ap p (2.11)

Substituting the pressure field ( )p r into equation (2.7), we find the normal displacement:

2 2*

1 3( ) 2 ,

4mpw r a ra

n pm

(2.12)

with the equivalent shear modulus *m defined as:

1 2

*1 2

1 1 1.

n nm m m

(2.13)

Using above the equation and substituting equations (2.5) and (2.12) into equation (2.6)

one can obtain the following relationship within the contact area:

Page 51: Time-Dependent Mechanics of Living Cells

51

2

2 2* *

32

16 2mp ra r

a Rp

dm

(2.14)

from which the mutual approach of the distant points in the two solids can be derived:

*

3.

8mp ap

dm

(2.15)

All parameters can be expressed in terms of total load F as in practice this parameter is

specified:

*

3 * * 3/2*

8 8,

3 3F a R

Rm

m d (2.16)

where *a R d . One special case of the Hertz contact problem which is of interest to

us is where one sphere is rigid ( 1m ) and of finite radius 1R and the other sphere is

elastic 2m but 2R . This is the case of indentation of a semi-infinite domain and for

this case the force-indentation relationship is

2 3/21

2

8.

3(1 )F R

md

n (2.17)

2.2 Linear viscoelasticity

A viscoelastic material undergoing deformation simultaneously stores and dissipates

mechanical energy. Viscoelasticity is the phenomenological theory which describes the

time-dependent response of such material. In this framework regardless of the

microstructure of the material, the relaxation processes are represented by exponential or

power law functions of time. In models based on exponential relaxation processes, stress

or strain functions can be represented by combining spring and dashpot elements. In such

models, the resultant exponential functions can have a single relaxation time if there is

only one dashpot element or a distribution of relaxation times if there are several. On the

other hand, another type of relaxation model that cannot be represented by simple

mechanical analogues but is encountered experimentally widely in the dynamics of

complex materials is power law relaxation.

Page 52: Time-Dependent Mechanics of Living Cells

52

2.2.1 Storage and loss moduli

The stress-strain relationship for a linear viscoelastic isotropic material can be expressed

in the time domain as:

ε ε ε

σd d d

2 ( ) d ( ){tr } d ( ) d ,d d d

t t t

t t G tam t t l t t t tt t t

I (2.18)

where εd /dt is the shear rate and ( )G ta t is the relaxation modulus designating the

deviatoric and the dilational part of the stress-strain relationship by 1a and 2a

respectively [111]. This equation implies that small changes in strain can be integrated

over time to yield the total stress.

The response of a material under oscillatory excitations is a typical way of measuring

shear viscoelastic properties. Let us assume that the material is subjected to an oscillatory

strain with amplitude ε0 and frequency w . Substituting a strain of the form

ε ε0( ) sint tw (with the corresponding strain rate of ε ε0d /dt= cos tw w ) into equation

(2.18) and performing some algebra gives the stress-strain relationship:

σ ε ε

εε ε ε

0 0

0 0

0 0

( ) ( )sin d sin ( )cos d cos

dsin cos ,

dt

t G t G t

GG t G t G

w t wt t w w t wt t w

w ww

(2.19)

where 0

( )sin dG Gw t wt t is the shear storage modulus and

0( )cos dG Gw t wt t the shear loss modulus. The physical meaning of the storage

and loss moduli can be explained as follows: if the material response is purely elastic then

the storage modulus G is the non-vanishing term in equation (2.19). In this case G

appears as the frequency-dependent shear modulus similar to a simple elastic material

where the stress is in phase with the strain: σ ε~ m . On the other hand, for a purely

viscous material 0G and the loss modulus G is the only non-vanishing term. This

results in a stress-response that has a phase lag of /2p with strain and /G w acts as the

Page 53: Time-Dependent Mechanics of Living Cells

53

frequency-dependent dynamic viscosity (similar to Newtonian fluid behavior

σ ε~ d /dth , where h is the viscosity). Unlike purely elastic or purely viscous materials,

viscoelastic materials exhibit a mixed behaviour with a phase lag of j (0 /2j p )

between the applied strain and the resultant stress.

From an experimental point of view, one way to determine the viscoelastic properties of a

material is to measure the stress response of the material in response to an applied

oscillatory strain of the form 0( ) sint te e w . Let us consider the measured stress-

response to have the form 0( ) sint ts s w j where 0s is the amplitude of stress

oscillations. Using equation (2.19), the following relationships can be used to calculate

the frequency-dependent storage and loss moduli of the material:

0

00

0

cos ,

sin .

G

G

sj

es

je

(2.20)

2.2.2 Standard linear solid model

In this section the standard linear solid model (also known as the Zener model) which has

been widely used to describe the time-dependent behaviours of biomaterials such as

living cells [51, 53, 54] (see Figure ‎2-2A) is presented. The differential equation

describing this model is:

1 1 2 1 2

d d.

dt dtK K K K K

s eh s h e (2.21)

Applying a sudden constant strain 0 ( )H te e (where ( )H t is the Heaviside step

function) or a sudden constant stress 0 ( )H ts s on a Zener viscoelastic material results

in a stress-relaxation or creep response :

1

0 2 1 ,Kt

K K e hs e Figure 2-2B, (2.22)

Page 54: Time-Dependent Mechanics of Living Cells

54

1 2

1 20 1

2 1 2

1 ,K K

tK K

Ke

K K Kh

se Figure 2-2C. (2.23)

Performing some mathematical substitutions, equation (2.21) can be solved to obtain the

dynamic response of a Zener material to an oscillatory load with G and G taking the

form (as plotted in Figure ‎2-2D):

2 2 21 2 2 1

2 2 21

21

2 2 21

( ),

K K K KG

KK

GK

h wh w

hwh w

(2.24)

2.2.3 Power law models

One of the most surprising features of a wide range of soft materials is that they display a

power law behaviour in which their relaxation spectrum lacks any characteristic time

scale. It is surprising because materials (including living cells) with very different

microstructural organization exhibit the same empirical behaviour. The creep and stress

relaxation responses of such power-law material can be written in the form:

0 0

0

0 0 '0

( ) /

( )

tt K

tt

t Kt

b

b

e s

s e (2.25)

where 0K is the parameter characterizes the softness or compliance of the material and 0t

and '0t are the normalization timescales that can be chosen arbitrary [64]. In other words,

0K is the ratio of stress to unit strain measured at an arbitrary chosen time 0t [112]. The

timescale invariant behavior of power law functions comes from the fact that changing 0t

or '0t does not affect the value of b . This equation can describe linear elastic and

Newtonian viscous behavior for b = 0 and b = 1 respectively. Using this equation, non-

exponential relaxation processes can be described in a very economical way with a small

number of parameters. Considering equations (2.19) and (2.25), the response of the

Page 55: Time-Dependent Mechanics of Living Cells

55

material to dynamic loading also follow a power law with the same exponent b ,

'', ~G G bw [113, 114].

Figure ‎2-2 The standard linear solid viscoelastic model.

(A) The standard linear solid model consists of a spring 2K in parallel with a spring 1K in series with a

dashpot h . (B) Stress-relaxation: the stress relaxes exponentially over time in response to application of a

sudden constant strain. (C) Creep: in response to application of a sudden constant stress, the strain increases

instantaneously until it reaches its final value. (D) Dynamic response of Zener model (with 1 1K pa,

2 0.1K pa and 1h pa.s): plot of the storage and loss moduli as a function of applied frequency.

Page 56: Time-Dependent Mechanics of Living Cells

56

2.3 Poroelasticity

The theory of poroelasticity studies the behaviour of a porous medium consisting of an

elastic solid matrix infiltrated by an interconnected network of fluid saturated pores.

2.3.1 Theory of linear isotropic poroelasticity

In this section the governing equations for a linear isotropic poroelastic material are

presented [115]. The constitutive equations are an extension of linear elasticity to

poroelastic materials first introduced by Biot [116]. Alternatively one can use the biphasic

model [117] that has been extensively applied in modelling the mechanics of articular

cartilage and other soft hydrated tissues. Biot’s formulation can be simplified when

poroelastic parameters assume their limiting values. Under the “incompressible

constituents” condition, the material exhibits its strongest poroelastic effect and the Biot

poroelastic theory can be mathematically transformed to the biphasic model.

Constitutive law Let us consider the quasistatic deformation of an isotropic fully

saturated poroelastic medium with a constant porosity. The constitutive equation relates

the total stress tensor σ to the infinitesimal strain tensor ε of the solid phase and the pore

fluid pressure p :

σ ε ε -2

2 tr( ) ,(1 2 )

s s

ss

GG p

n

nI I (2.26)

where sG and sn are the shear modulus and the Poisson ratio of the drained network

respectively, and εtr ( )q the variation in fluid content. This equation is similar to the

constitutive governing equation for conventional single phase linear elastic materials.

However the time dependent properties are incorporated through the pressure term that

acts as an additional external force on the solid phase.

Equilibrium equation In the absence of body forces and neglecting the inertial terms,

the local stress balance results in the equilibrium equation

Page 57: Time-Dependent Mechanics of Living Cells

57

σdiv 0. (2.27)

Darcy transport law Darcy’s law considers fluid transport through the non-

deformable porous medium

K pq (2.28)

where q is the filtration velocity and K the hydraulic permeability.

Continuity equation Neglecting any source density, the mass conservation of a fluid

yields:

div .tq

q (2.29)

2.3.2 Diffusion equation

Combining the continuity equation and Darcy’s law results in

2 .K pt

q (2.30)

Applying the equilibrium condition to the constitutive law, one can obtain the Navier

equations

2 div 0,(1 2 )

ss

s

GG p

nuu + (2.31)

where u is the vector of solid displacement for small deformations

ε 0.5 Tu u , and 2 designates the Laplacian operator.

The diffusion equation for εtr ( )q is obtained by combining equations (2.30) and

(2.31)

2PDt

qq (2.32)

where pD is the poroelastic diffusion coefficient:

Page 58: Time-Dependent Mechanics of Living Cells

58

2 (1 )(1 2 )s s

ps

GD K

nn

(2.33)

Equation (2.32) is the fundamental equation in the theory of consolidation which deals

with the time dependent settlement of porous materials. Derivation of the diffusion

equation implies that under the assumptions made in this section, there are three

independent sets of parameters ( sG , sn and K ) that characterize the mechanical

properties of a poroelastic medium.

2.3.3 Scaling law between poroelastic properties and microstructural parameters

One important consequence of considering a poroelastic cytoplasm is that the measured

macroscopic mechanical properties of the cell can be related to some coarse-grained

cellular microstructural parameters and the hydraulic pore size x of the cytoplasm can be

determined as a first step to understanding the microstructure of the cell. Considering a

simple Poiseuille flow inside a tube of radius r , the relationship between the flow rate Q

and the pressure gradient p is given by

4

,8r

Q pp

m (2.34)

where m is the viscosity of the fluid. For a porous body that contains n straight tubes per

unit area with porosity 2n rj p the filtration velocity is

4 2

.8 8n r r

q nQ p pp jm m

(2.35)

Taking into account irregularities, interconnectivities and tortuosities of the pores in a real

porous matrix the above relationship reads:

2

,4r

q pjmk

(2.36)

where k is the Kozney constant [118]. Considering equation (2.36) and a simple analogy

between the described Poiseuille flow and the flow through the porous matrix with an

Page 59: Time-Dependent Mechanics of Living Cells

59

average pore size x (Darcy’s law) leads to the following relationship for the hydraulic

permeability K

2

,4

Kj xk m

(2.37)

Substituting this equivalent expression for the hydraulic permeability into equation (2.33)

results in:

2(1 ).

(1 )(1 2 ) 4s s

ps s

ED

n j xn n k m

(2.38)

As a first approximation all of the parameters inside the parenthesis are assumed to be a

constant a and the functional dependence of all parameters with respect to the porosity

of the structure j is neglected. Therefore a fundamental scaling law for poroelastic

cytoplasm takes the form:

2

~ ,p

ED

xm

(2.39)

where m is interpreted as the interstitial fluid viscosity, and 2 1s sE G n the average

elasticity of the constituent solid network.

2.3.4 Poroelastic swelling/shrinking of gels under osmotic stress

The fundamental poroelastic equations for swelling/shrinking of gels under osmotic

perturbation are very similar to the ones presented in section 2.3.1 with a redefinition of

some parameters such as pore pressure and fluid flux [119]. Let us consider the initial

state of the gel to be homogeneous and free from mechanical load with 0C the initial

concentration of solvent inside the gel and 0m the initial chemical potential of the gel.

Equation (2.29) rewritten in terms of the concentration of the solvent in the gel is:

div .Ct

q (2.40)

Darcy’s law defines the migration of the solvent in the gel, thus equation (2.28)

transforms to

Page 60: Time-Dependent Mechanics of Living Cells

60

2

,K

mq (2.41)

where is the volume per solvent molecule and m the time-dependent chemical

potential of the solvent. Incompressibility of the network and the solvent imply that the

change in volume of the gel is only due to migration of solvent molecules into or out of

the network which changes the volume of solvent inside the gel:

ε 0tr( ) ( ).C C (2.42)

At thermodynamic equilibrium, the work done on each element of the gel equals the

change in free energy Wd written as σ ε 0tr( ) ( )W Cd d m m d. , where W is the

Helmholtz free energy per unit volume of the gel. In this relationship σ ε( . )tr d is the

mechanical work due to stress and 0( ) Cm m d is the work done by the chemical

potential. Using equation (2.42) one can write the change in free energy in terms of only

the strain tensor σ ε ε0tr( ) ( )tr /Wd d m m. which yields:

ε

σε

0( ) ( ).

W m mI (2.43)

The free energy can be written as a function of the strain tensor in the linear case of an

isotropic gel:

ε ε2 2[tr( ) ( ) ].(1 2 )

s

ss

W G trn

n (2.44)

Combining equations (2.43) and (2.44) results in equation (2.26) replacing pore pressure

by the term 0( )/m m :

σ ε ε - 02

2 tr .(1 2 )

s s

ss

GG

n m mn

I I (2.45)

Navier and diffusion equations similar to the Biot poroelasticity equations can then be

derived using the above equations:

21

div 0,(1 2 )

ss

s

GG m

nuu + (2.46)

Page 61: Time-Dependent Mechanics of Living Cells

61

2 .P

CD C

t (2.47)

These equations can be used to define the dynamics of gel swelling/shrinkage in response

to a change in extracellular osmolarity. Experiments in Chapter 4.2 examines this

situation in conjunction with the measurement of surface displacements of the gel by

defocusing microscopy to determine PD .

2.4 Basic 1D solutions for fundamental poroelasticity problems

2.4.1 Uniaxial strain (confined compression)

Let us consider the situation where all the components of the strain tensor except

/xx xu xe are zero so that the poroelastic material is only allowed to deform in one

direction. In this case the non-zero components of the stress tensor reduce to

-

-

2 (1 )

(1 2 )2 (1 2 )

.(1 2 ) (1 ) (1 )

s s

xx xxs

s s s syy zz xx xx

s s s

Gp

Gp p

ns e

nn n n

s s e sn n n

(2.48)

Substituting xu as the only component of the displacement field into the Navier equation

(2.31), reads

2

2

2 (1 )0

(1 2 )s s x

s

G u px x

nn

(2.49)

Combining equations (2.30) and (2.48) and using the fact that xxe q , the diffusion

equation for the pore pressure is obtained:

2

2xx

p

p pD

t x ts

(2.50)

Contrary to equation (2.32), equation (2.50) is an inhomogeneous diffusion equation that

can be solved by specifying a stress condition. In the following the one dimensional time-

Page 62: Time-Dependent Mechanics of Living Cells

62

dependent settlement of a poroelastic material subjected to certain types of boundary

conditions is studied.

2.4.2 One dimensional consolidation

Finite domain creep problem Consider a porous plunger that compresses a poroelastic

material confined in a cylinder of height L as shown in Figure ‎2-3. At 0t a certain

constant load 0(0, ) ( )xx t H ts s is applied on the solid matrix and held constant. Under

this condition the pressure diffusion equation (2.50) transforms into a homogeneous

diffusion equation

2

2.p

p pD

t x (2.51)

Figure ‎2-3 Compression of a porous solid with an interstitial fluid by a porous plunger.

The fluid diffuses out through the meshwork and exits through the plunger (red arrows) applying the external

load.

Page 63: Time-Dependent Mechanics of Living Cells

63

Assuming an ideal permeable plunger and zero flux at the bottom of the cylinder yields

0p at 0x and / 0p x at x L respectively. Considering the above

boundary and initial conditions one can solve equation (2.51) analytically:

2 20

1,3,..

4( , ) sin exp

2n

n xp x t n

m Lp

s p tp

(2.52)

where 2/4pD t Lt is a dimensionless time. The solid phase displacement field is

calculated employing equation (2.49) and the following boundary conditions 0xu at

x L :

2 20 2 2

1,3,..

(1 2 ) 8( , ) cos 1 exp

2 (1 ) 2s

xs s n

L n xu x t n

G n Ln p

s p tn p

(2.53)

Semi-infinite domain creep problem Now I consider the case where the poroelastic

material inside the cylinder is infinitely extended in the x direction. In this case, all the

boundary and initial conditions are the same as for the finite domain problem except that

0xu when x which implies that the dissipations arising from solid fluid

interactions vanishes to zero at infinity, / 0p x . The solutions for the pore pressure

and displacement field for a creep problem of a semi-infinite poroelastic domain can be

written as

0( , ) erf2 p

xp x t

D ts (2.54)

2

40

(1 2 )( , ) 2 erfc .

2 (1 ) 2p

xs D t

xs s p

Dt xu x t e x

G D tn

sn p

(2.55)

Figure ‎2-4 illustrates the normalized pore pressure and displacement field as a function of

depth at different times for 1pD . At 0x The settlement on the surface is

0

(1 2 )(0, ) 2 .

2 (1 )s

xs s

Dtu t

Gn

sn p

(2.56)

Page 64: Time-Dependent Mechanics of Living Cells

64

The solutions for application of a sudden loading of the top surface by a fluid with

pressure 0p can be derived similarly assuming the following boundary conditions

0(0, ) ( )p t p H t and (0, ) 0xx ts . The displacement solutions are exactly similar to

equations (2.53) and (2.55) replacing 0s with 0p . However the pore pressure solutions

take the form

2 20

1,3,..

4( , ) 1 sin exp ,

2n

n xp x t p n

m Lp

p tp

(2.57)

0( , )2 p

xp x t p erfc

D t (2.58)

for finite and infinite domains respectively.

Figure ‎2-4 Creep response of a poroelastic half-space.

Pore pressure (left) and displacement (right) solutions for the semi-infinite domain creep problem. The curves

are plotted as a function of depth at different times for 1pD .

In our experimental setup, cells have a finite thickness, therefore it is important to

investigate the time/length scales at which the semi-infinite solutions are applicable to

finite domain problems. Considering equation (2.58), at distance x L , it takes a time t

larger than 20.13 / pL D for the pore pressure to reach 5% of its maximum value at the

loaded boundary. Thus at times less than 20.13 / pL D the pore pressure perturbation at

the boundary 0x weakly influences the pore pressure at x L . In other words, at

distances beyond 3 pD t the pore pressure is less than 5% of its maximum at the

Page 65: Time-Dependent Mechanics of Living Cells

65

boundary. In conclusion, at short times, perturbations applied to the surface boundary are

confined to small regions near the boundary and thus the solution for the confined

problem can be estimated from semi-infinite solutions.

Figure ‎2-5 Creep response of a finite domain.

Top: Settlement response of the surface following a sudden application of aves = 250 Pa for a poroelastic

material with sE = 1 kPa, sn = 0 and pD = 40 µm2s-1. The dashed line is the semi-infinite response and the

lines are the solutions of equation (2.53) for poroelastic materials of thicknesses ranging from h = 2 to 8 µm.

Bottom: The displacement error considering the semi-infinite solution as the exact solution for the finite

thickness domain.

Furthermore, in relation to real AFM indentation experiments, to investigate the time

scales at which the half-space approximation can be employed, representative values of

the physical parameters derived from experimental data were used to compare the surface

Page 66: Time-Dependent Mechanics of Living Cells

66

displacements ( 0x ) in equations (2.53) and (2.56) for finite and semi-infinite

domains. As will be described later in Chapter 4, application of F ~ 6 nN via a spherical

tip with radius R = 7.5 µm results in indentation depths of 0d ~ 1 µm which leads to an

average stress of /ave F Rs p d ~ 250 Pa. Figure ‎2-5 shows the displacement as a

function of time considering thicknesses H = 2 - 8 µm for a poroelastic material. It can

be observed from this figure that for very small thicknesses (H < 3 µm) the semi-infinite

solution is applicable only at very short times ( t < 100 ms). For thicknesses of H ~ 5

µm (as shall be seen later this is approximately the height of the cells studied in this

thesis) the semi-infinite approximation is valid for the first ~ 300 ms of the process. The

half-space solution is a very good approximation for longer times ( t ~ 1 s) if the

thickness is H > 8 µm.

Semi-infinite domain stress relaxation Let us consider again the semi-infinite

problem but with sudden application of displacement (0, ) ( )u t UH t instead of load on

the top surface of a poroelastic material. This is the case of stress relaxation where the

pressure is suddenly increased due to fast compression of the top layer followed by

relaxation over time. Considering the far field conditions / 0u p x at x ,

the strain diffusion equation (2.32) transforms into a displacement diffusion equation

2

2.

x x

P

u uD

t x (2.59)

The displacement field satisfying the boundary conditions can be found as

( , ) erfc .2x

p

xu x t U

D t (2.60)

The corresponding pressure field becomes

2

4( , ) 1 p

xp D tDU

p x t eK tp

(2.61)

Page 67: Time-Dependent Mechanics of Living Cells

67

In Figure ‎2-6 the normalized pore pressure and the displacement field are plotted as a

function of depth at different times for 1pD .

Figure ‎2-6 Stress relaxation response of a poroelastic half-space.

Pore pressure (left) and displacement (right) solutions from the semi-infinite domain stress relaxation

problem. The curves are plotted as a function of depth at different times for 1pD .

2.5 Indentation of a poroelastic material

Conventional engineering methods such as compression, tensile and shear tests have

limitations for the characterisation of very soft materials due to difficulties in

manipulating soft and very small samples. Furthermore these methods do not have the

necessary accuracy and sensitivity to study the time-dependent dynamics of very soft

materials with short time scale responses. One of the best methods for characterisation of

very soft materials, especially cells, is an indentation test and as explained later in chapter

3, I conducted AFM dynamic micro-indentation experiments to study cell rheology in the

framework of poroelasticity. For indentation of a purely elastic material, there presently

exist several closed-form formulations such as Hertz’s equation (2.17) for obtaining

cellular elastic properties such as the shear modulus. Some closed-form expressions have

also been proposed for characterisation of time-dependent viscoelastic materials such as

indentation of a Zener viscoelastic material as presented in Chapter 3.7.3 . However due

to the higher complexity of the poroelastic equations, geometry and boundary conditions,

Page 68: Time-Dependent Mechanics of Living Cells

68

there exists no simple closed-form analytical solution that can be used to fit experimental

data. Therefore in this work an empirical expression derived from finite element (FE)

simulations [120] was employed to fit experimental poroelastic relaxation responses in

AFM indentation tests.

Figure ‎2-7 Spherical indentation and force relaxation a poroelastic material of finite thickness.

The indentor is pressed into the material until it reaches a target indentation depth of d . The force required to

induce such an indentation depth is iF applied with a short rise-time of rt . The applied force iF relaxes to

fF with the poroelastic characteristic time of t . (B) Compression of the solid matrix of a poroelastic

material with a short rise time rt t , pressurise the interstitial fluid. This pressurisation induces fluid

permeation expanding in a semi-spherical shape. The fluid migration can be affected by the substrate and for

h a the shape of expansion changes from a semi-spheroid to a 2D tube. (C) Experimental force-

relaxation data (grey dots) fitted with empirical solutions for force-relaxation of poroelastic materials

assuming the cell has a finite height (black) or that the cell can be approximated as a semi-infinite half-plane

(blue). Both empirical solutions fitted the experimental data well ( 2r > 0.95).

Page 69: Time-Dependent Mechanics of Living Cells

69

2.5.1 Indentation of a thin layer of poroelastic material

Very recently, detailed finite-element simulations of spherical indentation of thin layers

of poroelastic material (Figure ‎2-7) have been reported [121]. Recalling the Hertz

equation (2.17) for indentation of a semi-infinite elastic material, the reactive force on the

indentor can be expressed in terms of shear modulus G , Poisson ratio n and indentation

depth d : 1/2 3/2(8/3) /(1 ).HertzF GR d n For indentation of a poroelastic material of

finite thickness, the stress field is influenced by the substrate and thus the above equation

can be corrected for the effects of the substrate:

1/2 3/28

/ ,3 1 p

GF R f R hd d

n (2.62)

where pf is solved numerically and takes the following empirical form [121]:

20.46 0.82 / 2.36 /

/0.46 /p

R h R hf R h

R h

d dd

d (2.63)

For infinitely fast indentations ( rt t ) the fluid does not have enough time to

permeate through deformed regions and the poroelastic material behaves as an

incompressible elastic material (undrained condition) with the force response of

1/2 3/216

/ .3i pF GR f R hd d At long time scales, when the interstitial fluid has fully

redistributed (fully drained condition), the force imposed by the indenter is balanced by

the stress in the elastic porous matrix only. Under this condition, the force applied for the

prescribed indentation is 1/2 3/28

/3 1f p

s

GF R f R hd d

n. Thus, the Poisson ratio of

the solid matrix sn determines the ratio of forces at short and long timescales

/ 2 1i f sF F n . In intermediate regimes, the fluid permeates through pores and

force-relaxation can be approximated in terms of stretched exponential functions:

( )

expf

i f

F t F

F Fbat (2.64)

Page 70: Time-Dependent Mechanics of Living Cells

70

where 2

pD t

at is the poroelastic characteristic time required for force to relax from iF

to fF , and a and b are functions estimated empirically from the simulations that

depend on the contact radius a Rd and the height h of the layer:

2 3

4

2 3

4 5

1.15 0.44( / ) 0.89( / ) 0.42( / )

0.06( / )

0.56 0.25( / ) 0.28( / ) 0.31( / )

0.1( / ) 0.01( / ) .

R h R h R h

R h

R h R h R h

R h R h

a d d d

d

b d d d

d d

(2.65)

In my experiments, a typical height of h ~ 4.5 µm (at indention point) was measured for

HeLa cells and a and b were estimated from the above equations. Fitting experimental

force curves with the proposed stretched exponential function indicated that for h > 4.5

µm and d < 1.4 µm, approximating the cell to an infinite half-plane resulted in a less than

25% overestimation of the poroelastic diffusion constant pD (Figure ‎2-7C).

Page 71: Time-Dependent Mechanics of Living Cells

71

Chapter 3

Materials and Methods

Equation Chapter (Next) Section 1

3.1 Cell culture, generation of cell lines, transduction, and molecular

biology

In my experiments, I used three cell lines: cervical cancer HeLa cells, Madin-Darby

Canine Kidney epithelial (MDCK) and HT1080 fibrosarcoma cells. HeLa cells, HT1080

and MDCK cells were cultured at 37°C in an atmosphere of 5% CO2 in air in DMEM

(Gibco Life Technologies, Paisley, UK) supplemented with 10% FCS (Gibco Life

Technologies) and 1% PS. Cells were plated onto 50 mm glass bottomed Petri dishes

(Fluorodish, World Precision Instruments, Milton Keynes, UK). For MDCK cells, they

were cultured until forming a confluent monolayer. Prior to the experiment, the medium

was replaced with Leibovitz L-15 without phenol red (Gibco Life Technologies)

supplemented with 10% FCS.

To enable imaging of the cell membrane, we created a stable cell lines expressing the PH

domain of Phospholipase Cδ tagged with GFP (PHPLCδ-GFP), a phosphatidyl-inositol-

Page 72: Time-Dependent Mechanics of Living Cells

72

4,5-bisphophate binding protein that localises to the cell membrane. Briefly, PH-PLCδ-

GFP (a kind gift from Dr Tamas Balla, NIH) was excised from EGFP-N1 (Takara-

Clontech Europe, St Germain en Laye, France), inserted into the retroviral vector

pLNCX2 (Takara-Clontech), and transfected into 293-GPG cells for packaging (a kind

gift from Prof Daniel Ory, Washington University [122]). Retroviral supernatants were

then used to infect wild type HeLa cells, cells were selected in the presence of 1 mg.ml-1

G418 (Merck Biosciences UK, Nottingham, UK) for 2 weeks, and subcloned to obtain a

monoclonal cell line. Using similar methods, we created cell lines stably expressing

cytoplasmic GFP for cell volume estimation, GFP-actin or Life-act Ruby ([123]; a kind

gift of Dr Roland Wedlich-Soldner, MPI-Martinsried, Germany) for examination of the

F-actin cytoskeleton, GFP-tubulin for examination of the microtubule cytoskeleton, and

GFP-Keratin 18 (a kind gift of Dr Rudolf Leube, University of Aachen, Germany) for

visualisation of the intermediate filament network. HT1080 cells expressing mCherry-

LifeAct and MDCK cells expressing PHPLCδ-GFP were generated using similar

methods. The EGFP-10x plasmid was described in [124] and obtained through Euroscarf

(Frankfurt, Germany). Cells were transfected with cDNA using lipofectamine 2000

according to the manufacturer’s instructions the night before experimentation*.

3.2 Fluorescence and confocal imaging

In some experiments, we acquired fluorescence images of the cells being examined by

AFM using an IX-71 microscope interfaced to the AFM head equipped with an EMCCD

camera (Orca-ER, Hamamatsu, Germany) and piloted using µManager (Micromanager,

Palo-Alto, CA). Fluorophores were excited with epifluorescence and the appropriate filter

sets and images were acquired with a 40x dry objective (NA = 0.7). For staining of the

nucleus, cells were incubated with Hoechst 34332 (1 µg/ml for 5 min, Merck-

Biosciences).

*The cell lines described in this paragraph were generated with kind help from Dr. Charras.

Page 73: Time-Dependent Mechanics of Living Cells

73

In some cases, an AFM interfaced with a confocal laser scanning microscope (FV1000,

Olympus) was utilized to image the cellular indentation in the zx-plane. Images were

acquired with a 100x oil immersion objective lens (NA = 1.4, Olympus). Latex beads

attached to AFM cantilevers were imaged by exciting with a 647 nm laser and collecting

light at 680 nm. GFP tagged proteins were excited with a 488 nm laser and light was

collected at 525 nm. zx-confocal images passing through the centre of the bead were

acquired with 0.2 µm steps in z to give a side view of the cell before and after indentation

(Figure ‎3-2D).

3.3 Cell volume measurements

To measure changes in HeLa cell volume in response to osmotic shock, confocal stacks

of cells expressing cytoplasmic GFP were acquired at 2 min intervals using a spinning

disk confocal microscope (Yokogawa CSU-22, Yokogawa, Japan) with 100x oil

immersion objective lens (NA = 1.4, Olympus) and a piezo-electric z-drive (NanoscanZ,

Prior, Scientific, Rockland, MA). Stacks consisted of 40 images separated by 0.2 µm and

were acquired every 2 min for a total of 30 mins. Exposure time and laser intensity were

set to minimize photobleaching. A custom written code in Matlab (Mathworks Inc,

Cambridge, UK) was used to process the z-stacks and measure the cell volume at each

time step. Briefly, the background noise of stack images was removed, images were

smoothed, and binarised using Matlab Image Processing Toolbox functions. Following

binarisation, series of erosion and dilatation operations were performed to create a

contiguous cell volume image devoid of isolated pixels (Figure ‎3-1A, B). The sum of

nonzero pixels in each stack was multiplied by the volume of a voxel to give a measure of

cell volume at each time step. All experiments followed the same protocol: five stacks

were captured prior to change in osmolarity and then cell volume was followed for a

further 25 minutes (see Figure ‎5-2A).

Page 74: Time-Dependent Mechanics of Living Cells

74

Figure ‎3-1 Cell volume measurement.

(A, B) The volume of HeLa cells was measured by acquiring fast xyz-t confocal stacks. (A) Image of HeLa

cytoplasmic GFP cell showing the bottom of the cell before (I) and after (II) application of hyperosmotic

shock induced by PEG-400. The decrease in cell volume increases the concentration of GFP molecules

resulting in an increase in fluorescence intensity. (B) Reconstruction of cell profile from confocal z-stacks

before I and after II application of PEG-400. (C-I) xy image of MDCK monolayer expressing PH-PLCδ. (C-

II) zx section of MDCK cells expressing a membrane marker (PH-PLCδ, green) and stained for nucleic acids

with Hoechst 34332 (blue). Nuclei localised to the basal side of cells far from the apex. (C-III) zx profile of

MDCK cells before (green) and after (red) application of 500 mM sucrose. Cell height decreased significantly

in response to increase in medium osmolarity. Scale bars = 10µm.

The volume of PH-PLCδ MDCK cells in response to osmotic perturbations was estimated

by measuring the heights of cell from confocal z-slice images after and before application

osmotic shocks (Figure ‎3-1C). During the osmotic perturbations, I verified that the xy

projected area of the cells and the xy positions of cell-cell junctions were not perturbed

significantly and therefore measuring the change in cell height was a good approximation

for estimating changes in cell volume.

Changes in extracellular osmolarity were effected by adding a small volume of

concentrated sucrose, 400-Dalton polyethylene glycol (PEG-400, Sigma-Aldrich, [125]),

Page 75: Time-Dependent Mechanics of Living Cells

75

or water to the imaging medium. When increasing osmolarity by addition of osmolyte,

MDCK cells were treated with EIPA (50 µM, Sigma-Aldrich), an inhibitor of regulatory

volume increases [126]. When decreasing the osmolarity by addition of water, the cells

were simultaneously treated with the inhibitors of regulatory volume decreases, NPPB

(200 μM; Tocris, Bristol, UK) and DCPIB (50 µM, Tocris), to achieve a sustained

volume increase [126]. Cells were incubated with these inhibitors for 30 minutes prior to

the addition of osmolytes (PEG-400, sucrose or water).

3.4 Disrupting the cytoskeleton

3.4.1 Pharmacological treatments

Cells were incubated in culture medium with the relevant concentration of drug for 30

min prior to measurement. The medium was then replaced with L-15 with 10% FCS plus

the same drug concentration such that the inhibitor was present at all times during

measurements. Cells were treated with latrunculin B (to depolymerise F-actin, Merck-

Biosciences), nocodazole (to depolymerise microtubules, Merck-Biosciences), and

paclitaxel (to stabilize microtubules, Merck-Biosciences), and blebbistatin (to inhibit

myosin II ATPase, Merck-Biosciences).

3.4.2 Genetic treatments

To examine the effect of uncontrolled polymerization of cytoplasmic F-actin, Dr.

Moulding transduced HeLa cells stably expressing Life-act ruby with lentivirus encoding

WASp I294T as described in [26]. Lentiviral vectors expressing enhanced GFP fused to

human WASp with the I294T mutation were prepared in the pHR’SIN-cPPT-CE and

pHR’SIN-cPPT-SE lentiviral backbones as described previously [26, 127]. Lentivirus

was added to cells at multiplicity of infection of 10 to achieve approximately 90%

transduction.

Page 76: Time-Dependent Mechanics of Living Cells

76

To examine the perturbations to cytoskeletal proteins, Dr. Charras generated the plasmids

described below, and I purified the cDNA and transfected the cells. To examine the effect

of uncontrolled polymerisation of tubulin, I transfected HeLa cells with a plasmid

encoding γ-tubulin-mCherry, a microtubule nucleator [128, 129]. To disrupt the keratin

network of HeLa cells, keratin 14 R125C-YFP (a kind gift from Prof Thomas Magin,

University of Leipzig) was overexpressed, a construct that acts as a dominant mutant and

results in aggregation of endogenous keratins [130]. To disrupt F-actin crosslinking by

endogenous α-actinin, cells were transfected with a deletion mutant of α-actinin lacking

an actin-binding domain (ΔABD-α-actinin, a kind gift of Dr Murata-Hori, Temasek Life

Sciences laboratory, Singapore). Cells were transfected with cDNA using lipofectamine

2000 the night before experimentation.

3.5 Visualising cytoplasmic F-actin

To visualise cytoplasmic F-actin density, cells were fixed for 15 minutes with 4% PFA at

room temperature, permeabilised with 0.1% Triton-X on ice for 5 min, and passivated by

incubation with phosphate buffered saline (PBS) and 10 mg/ml bovine serum albumin

(BSA) for 10 min. They were then stained with Rhodamine-Phalloidin (Invitrogen) for 30

min at room temperature, washed several times with PBS-BSA, and mounted for

microscopy examination on a confocal microscope.

Page 77: Time-Dependent Mechanics of Living Cells

77

Figure ‎3-2 AFM experimental setup.

(A) Schematic diagram of the experiment. (A-I) The AFM cantilever is lowered towards the cell surface with

a high approach velocity approachV ~ 30 µm.s-1. (A-II) Upon contacting the cell surface, the cantilever bends

and the bead starts indenting the cytoplasm. Once the target force MF is reached, the movement of the

piezoelectric ceramic is stopped at MZ . The bending of the cantilever reaches its maximum. This rapid force

application causes a sudden increase in the local stress and pressure. (A-III) and (A-IV) Over time the cytosol

in the indented area redistributes inside the cell and the pore pressure dissipates. Strain resulting from the

local application of force propagates through the elastic meshwork and at equilibrium; the applied force is

entirely balanced by cellular elasticity. Indentation (I-II) allows the estimation of elastic properties and

relaxation (III-IV) allows for estimation of the time-dependent mechanical properties. In all panels, the red

line shows the light path of the laser reflected on the cantilever, red arrows show the change in direction of

the laser beam, black arrows show the direction of bending of the cantilever, and the small dots represent the

propagation of strain within the cell. (B) Temporal evolution of the cantilever position and force applied

during the experiment. Before the cantilever contacts the cell, no force is applied (I). After contact, force

increases until the target force (II). Finally, force relaxes (III) until equilibrium (IV). (C) Schematic drawing

showing the zone of indentation. A bead of radius R is pressed into the cell surface creating a depression of

radius a and depth ad . The compression of the cell surface by the bead induces a pressure increase that

squeezes a volume 2~ ad of fluid out of the indentation region (arrows). (D) Z-x confocal image of a HeLa

cell expressing PH-PLCδ1-GFP (a membrane marker) corresponding to phases (I) and (IV) of the experiment

described in A. The fluorescent bead attached to the cantilever is shown in blue and the cell membrane is

shown in green. Scale bar = 10 μm.

Page 78: Time-Dependent Mechanics of Living Cells

78

3.6 Atomic force microscopy indentation experiments

Force-distance and force-relaxation measurements were acquired with a JPK

Nanowizard-I (JPK instruments, Berlin, Germany) interfaced to an inverted optical

fluorescence microscope (IX-81 or IX-71, Olympus). For most experiments, AFM

cantilevers (MLCT, Bruker, Karlsruhe, Germany) were modified by gluing beads to the

cantilever underside with UV curing glue (UV curing, Loctite, UK). Cantilever spring

constants were determined prior to gluing the beads using the thermal noise method

implemented in the AFM software (JPK SPM, JPK instruments). Prior to any cellular

indentation tests, the sensitivity of the cantilever was set by measuring the slope of force-

distance curves acquired on glass regions of the petri dish. For measurements on cells, I

used cantilevers with nominal spring constants of 0.01 N.m-1

and fluorescent latex beads

with radii of 7.5 μm (ex645/em680, Invitrogen). For measurements on hydrogels, I used

cantilevers with nominal spring constants of 0.6 N.m-1

and glass beads with radii of 25

μm (Sigma).

During AFM experiments, the bead on the cantilever was aligned over regions near the

cell nucleus and measurements were acquired in several locations in the cytoplasm

avoiding the nucleus (except when doing spatial mapping, see section 3.7.5). To

maximize the amplitude of stress relaxation, the cantilever tip was brought into contact

with the cells using a fast approach speed ( approachV > 10 µm.s-1

) until reaching a target

force MF (Figure ‎3-2A-I, A-II, B). With these settings, force was applied onto the cells in

less than 35 ms on average, which is fast compared to the expected poroelastic relaxation

time (~0.1-10 s, see Chapter 4 section 1.1). Hence, in the analysis of stress relaxation I

assumed that force application was quasi-instantaneous. Upon reaching the target force

MF the piezoelectric ceramic length was kept constant at MZ and the force-relaxation

curves were acquired at constant MZ (Figure ‎3-2A-III, A-IV). After 10 s, the AFM tip

was retracted with the same speed as the approach.

Page 79: Time-Dependent Mechanics of Living Cells

79

During force-relaxation measurements, the z-closed loop feedback implemented on the

JPK Nanowizard was used to maintain a constant z-piezo height. When acquiring force-

relaxation curves with high approach velocities ( approachV > 10 µm.s-1

), the first 5 ms of

force-relaxation were not considered in the data analysis due to the presence of small

oscillations in z-piezoelectric ceramic height immediately after contact due to the PI

feedback loop implemented in the JPK software. When acquiring force-relaxations curves

for averaging over multiple cells, I used approach velocities approachV ~ 10 µm.s-1

that did

not give rise to oscillations in length of the z-piezoelectric ceramic.

3.6.1 Measuring indentation depth and cellular elastic modulus

The approach phase of the AFM force-distance curves (Figure ‎3-3) was analysed to

extract the bulk elastic modulus E , which is linearly related to the shear modulus G

through the Poisson ratio via equation (2.4); 2 (1 )E G n . During indentation (after

contact) the force the cantilever is related to the cell stiffness and the deflection of the

cantilever d by

0( ),c cF k d k d d (3.1)

where ck is the cantilever’s spring constant, d is the deflection of cantilever (correlated

to changes in reflection of AFM laser beam which is detected by photo-detector), and 0d

is the deflection offset at the point of contact where the force is zero.

The indentation depth d is calculated by subtracting the cantilever deflection d from

the piezo translation z

0 0

0 0 0

( ) ( )( ) ( ) ,

z d z z d dz d z d w w

d (3.2)

where 0z is the translation of the piezo at the contact point, ( )w z d and

0 0 0w z d are the transformed variables. E was then estimated by fitting the force-

indentation data with a Hertzian contact model between a sphere and an infinite half

Page 80: Time-Dependent Mechanics of Living Cells

80

space (equation (2.17)). In this framework, the relationship between the applied force F

and the indentation depth d is:

1/2 3/22

4.

3 1E

F R dn

(3.3)

with n the Poisson ratio, and R the radius of the spherical indenter. The above Hertz

model of indentation can be used to compute the elastic modulus of a linear elastic

material of infinite thickness (see Chapter 2.1.1). To limit the effects of finite cell

thickness, in my analysis I only considered the force-indentation/relaxation curves in

which the indentation depths were less than 25% of the cell height. This ensures errors of

less than 20% in the estimation of the elastic modulus (see equation 2.63) in my

experiments. To use the Hertz model in this work I assumed the cell to be a isotropic,

homogeneous and linear elastic materials and the contact surfaces were assumed to be

frictionless and non-adhesive [131]. Although these assumptions are not fully accurate for

indentation of adherent cells, the Hertz model is still widely used for characterisation of

cells and soft materials ‎providing fairly accurate measurements of elastic modulus [132].

Figure ‎3-3 AFM force-distance curve.

The approach phase of AFM force-distance curve (Figure ‎3-2A-I and II). The force distance AFM data

(dotted points) is fitted with two linear (blue line) and nonlinear (red line) curves for the noncontact and

contact portions of the curve respectively. The contact point is estimated by choosing the point that minimises

the total mean square error of the fitted curves.

Page 81: Time-Dependent Mechanics of Living Cells

81

Together with equations (3.1) and (3.2), contact mechanics models (like equation (3.3))

can be fit to AFM indentation data. However the main challenge in analysing AFM force-

distance curves is to design an automated routine to find the correct contact point which

enables us to compute the indentation depth. The contact point between the cell and the

AFM tip was estimated using the method described in [133] implemented in our

laboratory, see Figure ‎3-3. In brief, as a first step the force distance-curve is transformed

into a d -w curve introducing the variable ( )w z d . Then our custom written Matlab

code marches along all points of the d -w curve, and assumes each point is the potential

contact point. Considering ( 0iw , 0

id ) as the contact point, at each iteration i the initial

portion of the curve (assumed to be the noncontact region) was fitted with a line while the

latter portion (assumed to be the portion of the curve in contact) was fitted with a power

function as in equation (3.3):

3/20 0( ) ( ) ,i i id d b w w (3.4)

where ib is the sole fitting parameter. The total mean square error (MSE) from both fits

was calculated and the contact point ( 0nw , 0

nd ) was chosen as the point where the total

MSE reaches its minimum value at i n with 1/2 24 / 3(1 )ncb ER kn (see

Figure ‎3-3 ).

Comparing the long time-scale and short time-scale limits allows for estimation of the

Poisson ratio of the ‎solid matrix; (0)/ ( ) 2(1 )sF F n . However for cellular

indentations the force relaxation curves do not reach a relaxation plateau for estimation of

( )F and thus for estimation of the elastic modulus, I assumed a Poisson ratio of sn

=0.3.‎

3.6.2 Measurement of the poroelastic diffusion coefficient

A brief description of the governing equations of linear isotropic poroelasticity and the

relationship between the poroelastic diffusion constant pD , the shear modulus G , and

Page 82: Time-Dependent Mechanics of Living Cells

82

hydraulic permeability K are given in Chapter 2 section 3. To experimentally measure

the cellular poroelastic properties, I monitored stress-relaxation following rapid local

application of force by AFM microindentation (Figure ‎3-3, Figure ‎3-4 and Figure ‎5-1A). I

analysed my experiments as force-relaxation in response to a step displacement of the cell

surface. No closed form analytical solution for indentation of a poroelastic infinite half

space by a spherical indentor exists. However, an approximate solution obtained by FE

simulations gives [120]:

0.908 1.679( )

0.491e 0.509e ,f

i f

F t F

F Ft t (3.5)

where /( )pD t Rt d is the characteristic poroelastic time required for force to relax

from iF to fF . Cells have a limited thickness h and therefore the infinite half-plane

approximation is only valid at time-scales shorter than the time needed for fluid diffusion

through the cell thickness: 2~ /hp pt h D . As we shall see in the next chapters, in

experiments on HeLa cells, I measured h ~ 5 µm and pD ~ 40 µm2.s

-1 setting a time-

scale hpt ~ 0.6 s. I confirmed numerically that for times shorter than ~ 0.5 s,

approximating the cell to a half-plane gave errors (calculated by comparing the half space

approximation and the finite thickness solution) of less than 20% (Figure ‎2-5 and

Figure ‎2-7C). For short time-scales, both terms in equation (3.5) are comparable and

hence as a first approximation, the relaxation scales as ~e t . Equation (3.5) was utilized

to fit our experimental relaxation data, with pD as single fitting parameter and I fitted

only the first 0.5 s of relaxation curves to consider only the maximal amplitude of

poroelastic relaxation and minimise errors arising from finite cell thickness.

3.6.3 Determining the apparent cellular viscoelasticity

To determine the single phase viscoelastic parameters, I modelled cells using a standard

linear model consisting of a spring-damper (stiffness 1k and apparent viscosity h ) in

parallel with another spring (stiffness 2k ), as described in [53] (see Figure ‎2-2). In this

Page 83: Time-Dependent Mechanics of Living Cells

83

model, the applied force decays exponentially when the material is subjected to a step

displacement at t = 0:

1( ).

k tf

i f

F t Fe

F Fh (3.6)

The spring constant 1k scales with the elastic modulus 1 ~k E (estimated from approach

phase of indentation tests) and therefore the experimental force relaxation curves were

fitted using equation (3.6), with h as the sole fitting parameter.

3.6.4 Measuring cell height

To measure cell height at each contact point, I collected a force-distance curve on the

glass substrate next to the cell being examined. The 0z calculated from the force-distance

curve on the glass was used as a height reference. As shown in Figure ‎3-4, at each

indentation point, the cell height was estimated by subtracting the 0z of the cell from the

reference 0z .

Figure ‎3-4 Calculating the height of cell at contact point.

A force-distance curve acquired on the glass substrate next to the cell examined was used as the reference to

estimate the height of the cell at the contact point

Page 84: Time-Dependent Mechanics of Living Cells

84

3.6.5 Spatial probing of cells

To investigate spatial variations in the cellular poroelastic properties, I acquired

measurements at several locations on the cell surface along the long axis of the cell with a

target force of MF = 4 nN. To locate indentation points in optical images of the cell, a

fluorescence image of the bead resting on the glass surface close to the cell was acquired,

and the centre of the bead was used as a reference point. This reference point was used to

estimate the height of the cell for each measurement, as well as the position of indentation

on the cell surface, the xy-coordinates of each stress relaxation measurement being

known from the AFM software. The position of the measurement point relative to the

nucleus was estimated by acquiring a fluorescence image of the nucleus and calculating

the centre of mass of the nucleus (see Figure ‎3-5). This also enabled us to know whether

the measurement point was above the nucleus or not.

Page 85: Time-Dependent Mechanics of Living Cells

85

Figure ‎3-5 Spatial AFM measurement

(I) Representative phase contrast image of a HeLa cell and the location of AFM measurements on its surface.

The position of the height reference, where the AFM tip contacts the glass surface was also chosen as the xy

reference (fluorescence image of the bead, IV). The location of the nucleus was determined using Hoechst

34332 staining and the centroid of the nucleus (red dot in I) was used as a reference for displaying the

measured properties. To determine the exact pathline of the bead and to calculate the distance of each point

from the centre of nucleus, images of the cell (II), the cell nucleus (III), and the bead (IV) were overlaid (I).

Scale bars = 10 μm.

3.7 Microinjection and imaging of quantum dots

PEG-passivated quantum dots (qdots 705, Invitrogen) were diluted in injection buffer (50

mM potassium glutamate, 0.5 mM MgCl2, pH 7.0) to achieve a final concentration of 0.2

μM and microinjected into HeLa cells as described in [134]. Quantum dots were imaged

on a spinning disk confocal microscope by exciting at 488 nm and collecting emission

above 680 nm. To qualitatively visualize the extent of quantum dot movement, time

series were projected onto one plane using ImageJ.

Page 86: Time-Dependent Mechanics of Living Cells

86

3.8 Rescaling of force relaxation curves

Plotting the obtained experimental relaxation curves in different scale coordinates and/or

normalising them in different ways can provide useful insights into the functional nature

of these curves. For instance it is difficult to differentiate between exponential relaxation

curves and weak power law curves unless the curves are plotted in log-log coordinates.

Furthermore, to differentiate between poroelastic and viscoelastic responses during

indentation tests several methods for normalisation of relaxation curves have been

proposed [120, 135-137]. One successfully applied method is based on the collapse of

normalised force relaxation curves acquired with different indentation depths onto a

single master curve [137]. Prompted by this, I have proposed the following more general

force normalisation method:

( ) ( *)

,( 0) ( *)F t F t tF t F t t

(3.7)

where *t can be any arbitrary time. Let us consider that the force relaxation is decaying

function of the form 1 2( ) ( , , ,..., )nF t F t t t t that is a function of a number of (finite or

infinite) characteristic relaxation times it . For force relaxation curves acquired with

different indentation depths, if the characteristic relaxation times (or in general the decay

function F ) do not have any form of functional dependency on indentation depth (as

during linear viscoelastic or power law responses), force normalisation using equation

(3.7) results in the collapse of all the relaxation curves onto a single curve. In contrast, the

characteristic relaxation time in poroelastic materials depends on the indentation depth

and force-normalisations alone will not result in collapse of curves onto a single master

curve.

Considering the spherical indentation of a purely poroelastic material, the force relaxation

can be written in the form ( ) [ (0) ( )] ( / ) ( )pF t F F P D t R Fd where P is the

poroelastic response function (decaying in an exponential manner from 1 at 0t to 0 at

t ) and (0)F and ( )F are the forces at very short (t = 0+) and very long ‎(t = ∞)‎‎

Page 87: Time-Dependent Mechanics of Living Cells

87

timescales, respectively. For any given indentation depth d , using the normalisation in

equation (3.7) results in:

( / ) ( * / )( ) ( *)

( 0) ( *) 1 ( * / )p p

p

P D t R P D t RF t F t tF t F t t P D t R

d dd

(3.8)

To obtain collapse of force relaxation curves onto a single master curve for any *t using

the above equation time must be rescaled with indentation depth. First, *t must be

selected so that it is proportional to the ‎indentation depth d specific to each curve:

*t ad where a (with unit of s. μm-1

) is an arbitrary ‎number fixed for all curves. Then

rescaling of time with respect to indentation depth ( /t d ), leads to collapse of all

experimental curves onto one master curve.

Note that, if the chosen *t is such that * / pt R Dt d , then the force relaxation

curve reaches a plateau: ( ) ( )F t F and ( * / ) 0pP D t Rd and equation (3.7)

becomes:

( ) ( *)

( / ),( 0) ( *) p

F t F t tP D t R

F t F t td (3.9)

and therefore rescaling of time with respect to indentation depth ( /t d ), is sufficient to

obtain collapse of all relaxation curves as in [137].

3.9 Polyacrylamide hydrogels

For comparison with cells, I acquired force-relaxation measurements on polyacrylamide

(PAAm) hydrogels, which are well-characterised poroelastic materials. For the

measurements, we made gels of 15% acrylamide-bis-acrylamide crosslinked with

TEMED and ammonium persulfate following the manufacturer’s instructions (Bio-Rad,

UK).

Page 88: Time-Dependent Mechanics of Living Cells

88

3.10 Particle tracking using defocused images of fluorescent beads

Optical images of fluorescent beads acquired by wide-field microscopy can exhibit

different shapes depending on the distance between the bead and the plane of focus. The

in-focus image of a bead is a spot with a certain diameter, while when the plane of focus

is away from the bead centre by a distance of a few micrometres, sets of concentric rings

are visualised (see Figure ‎3-6A). The z-distance of the particle from the imaging plane is

precisely correlated to the radius of the outer ring. As a very simple 3D particle tracking

technique, calculating the centre and radius of the outer ring over several frames over

time can provide a very precise means for tracking the coordinates of a bead: ( , , , )x y z t

[138]. This technique is presented in Chapter 4.2 and was employed to study the

swelling/shrinking kinetics of cell (Figure ‎4-7).

3.10.1 Experimental protocol and calibration setup

Yellow-Green carboxylate-modified fluorescent microspheres 0.5 µm in diameter

(FluoSpheres, Molecular Probes, Invitrogen) were coated with collagen-I following the

protocol suggested by the manufacturer. To attach beads to the cell membrane prior to

experiments, Life-act ruby HeLa cells were incubated for ~ 30 min with a dilute solution

(100x dilution) containing fluorescent collagen coated beads.

A Piezo-electric z-stage (NanoscanZ, Prior, Scientific, Rockland, MA) was used to

calibrate defocused images and to estimate the height of the cell at the location of the

attached beads (Figure ‎3-6). The z-stage was fixed on top of the optical fluorescence

microscope stage (IX-71, Olympus) and was piloted using µManager (Micromanager,

Palo-Alto, CA) to step up/down while taking epifluorescence images. To find the

relationship between the z-positions of a bead and the radius of the outer rings formed in

the defocused images, fluorescence z-stack images of the bead with a step size of 100 nm

were acquired. To measure the poroelastic diffusion constant using equation (4.6), the

height of the bead also needed to be determined. For this, the z-position of the bead with

Page 89: Time-Dependent Mechanics of Living Cells

89

respect to the bottom of the cell was estimated by acquiring fluorescence z-stack images

of Lifeact-Ruby HeLa cells with step size of 200 nm over large distances (~ 10 μm,

starting from a few micrometres above the cell and ending a few micrometres below the

cell-glass interface). The height of the cell was measured as the distance between the

plane that shows an in-focus image of the stress fibres at the bottom of the cell and the

plane focused on the bead (Figure ‎3-6D).

Figure ‎3-6 High resolution defocusing microscopy of fluorescent bead.

(A) Images of a 500 nm fluorescent bead attached to the cell membrane taken at different distances from bead

centre appears as concentric rings. The radius of the outermost ring in the image can be related to the distance

between the bead centre and the plane of imaging. (B) The averaged radial intensity curves calculated when

the imaging plane is moved in 100 nm increments away from the bead focal plane. The radius of outer rings

can be estimated using these curves. Inset is the normalised averaged radial intensity curves. To determine the

radius of outer rings, the peak of each curve can be fitted with a Gaussian function. (C) The radial position of

the outer ring is linearly related to the distance between the imaging plane and the bead focal plane. Using

this concept the z-position of a particle can be tracked with very high accuracy (~ 10 nm). (D) The height of a

bead attached to the cell is estimated by focusing on the bottom of the dish in (I) where stress fibres are in

focus, and on the bead attached to the cell membrane in (II).

Page 90: Time-Dependent Mechanics of Living Cells

90

3.10.2 Radial projection

The 2D defocused images of a bead are circularly symmetric so a 1D radius function was

used to detect the radius of the outer ring in each defocused image. Since there were

several beads attached to the cell membrane, the rough central position of the bead of

interest was selected visually and then the precise coordinates of the centroid of the

chosen bead ( , )c cx y were estimated by fitting a 2-D Gaussian function to the image. An

arbitrary annulus interval was set to generate a radial distance vector

, 0...ir i i N . By knowing the precise centre of the image, the distance of each

pixel in the image from the centroid was estimated. According to the radial distance of

each pixel d , there are total of iP pixels lying between radii 1ir and 1ir . For each pixel

within the annulus between radii 1ir and 1ir the averaged radial intensity vector ( )R ir

(see Figure ‎3-6B) was calculated by splitting the intensity of each pixel I in this annulus

linearly according to its radial position d :

1

1

11

11

1( )

.

i

i

P

R i n nPn

nn

ii i

ni

i i

r w I

w

d rif r d r

w r dif r d r

(3.10)

To find an estimate with sub-pixel resolution ‎for the radial position of the outer rings, two

approaches were implemented. In the first approach the radius of the outer ring was

calculated by finding the position of the outer peak in the averaged radial intensity curve

(using an inbuilt Matlab function) and fitting a rotationally symmetric Gaussian function

to the intensity curve neighbouring this point. In the second approach the averaged radial

intensity curve was normalised to its maximum intensity as shown in the inset of

Figure ‎3-6B and the intersection of the curve with the line y = 0.05 provided an estimate

for the radial position of the ring where the intensity of the outer ring diminishes to zero.

Page 91: Time-Dependent Mechanics of Living Cells

91

3.11 FRAP experiments

I performed the fluorescence recovery after photobleaching (FRAP) experiments

described in the following with kind help from Marco Fritzsche. FRAP experiments were

performed using a 100x oil immersion objective lens (NA = 1.4, Olympus) on a scanning

laser confocal microscope (Olympus Fluoview FV1000; Olympus). Fluorophores

including GFP-tagged proteins and CMFDA (a small cytoplasmic fluorescent probe, 5-

chloromethylfluorescein diacetate, Celltracker Green, Invitrogen) were excited at a

wavelength of 488 nm. For experiments with CMFDA, cells were incubated with 2 µM

CMFDA for 45 min before replacing the medium with imaging medium. For experiments

with cytoplasmic GFP or EGFP-10x, cells were transfected with the plasmid of interest

the day before experimentation. To obtain a strong fluorescence signal and minimize

photobleaching, a circular region of interest (ROI, 1.4 µm in diameter) in the middle of

the cytoplasm was imaged, setting the laser power to 5% and 1% of the maximum output

(488nm wave length, nominal output of 20mW) for GFP and CMFDA, respectively. Each

FRAP experiment started with five image scans to normalise intensity, followed by a 1s

bleach pulse produced by scanning the 488nm laser beam line by line (at 100% power for

GFP and 25% power for CMFDA) over a circular bleach region of nominal radius nr =

0.5 μm centred in the middle of the imaging ROI. To sample the recovery sufficiently

fast, an imaging ROI of radius of 0.7 μm was chosen allowing acquisition of fluorescence

recovery with a frame rate of 50 ms per frame.

Page 92: Time-Dependent Mechanics of Living Cells

92

Figure ‎3-7 Estimation of the effective bleach radius in HeLa cells FRAP experiments.

(I) Confocal image of cells loaded with the cytoplasmic indicator CMFDA (green). Nuclei (blue) were

stained with Hoechst 34332. Scale bar = 10 µm. (II-III-IV) Time series of a photobleaching experiment in the

cytoplasm of the cells shown in (I, indicated by the white circle). The white circle in images (II-IV) denotes

the nominal photobleaching zone. In all three images, scale bars = 1 µm. (II) shows the last frame prior to

photobleaching. (III) shows the first frame after photobleaching. The bleached region is clearly apparent and

larger than the nominal photobleaching zone. (IV) After a long time interval, fluorescence intensity re-

homogenises due to diffusion. (V) Fluorescence intensity immediately after photobleaching in the cytoplasm

centred on the photobleaching zone in frame (III). Fluorescence was normalised to the intensity far away

from the photobleaching zone.

Page 93: Time-Dependent Mechanics of Living Cells

93

3.11.1 Effective bleach radius

Most experiments were carried out on cytoplasmic fluorophores. Because of their rapid

diffusion within the cytoplasm during the bleaching process, the effective radius over

which photobleaching takes place ( er ) is larger than the nominal radius ( nr ). In my

experiments to measure the translational diffusion coefficient TD precisely, I needed to

first determine er . To experimentally estimate the effective radius er , I performed a

separate series of FRAP experiments on cells loaded with CMFDA in hyperosmotic

conditions. An imaging area larger than the effective photobleaching radius (4 µm in

diameter) was empirically chosen and photobleaching recovery in this area was imaged

with a frame rate of 0.2 s/frame. Next, using a custom written Matlab program, I

calculated the radial projection profile of the fluorescence intensity in the imaging region

for the first postbleach image ( , 1)I r t and then fitted the experimental intensity profile

with the bleach equation for a Gaussian laser beam to determine er (Figure ‎3-7):

2

2( ) exp exp 2 ,

e

rI r K

r (3.11)

where r is the radial distance from the center of bleach spot and K is the bleaching

constant related to the intensity of the bleaching laser and the properties of the

fluorophore.

3.11.2 FRAP analysis

FRAP recovery curves ( )I t were obtained by normalising the mean fluorescence intensity

within the nominal bleach area for each frame to the average fluorescence intensity of the

first prebleach image. The translational diffusion coefficient TD was estimated using the

model for a uniform-disk laser profile and ideal bleach [139, 140] following the methods

described in [141]:

2 2 2

0 1( ) exp ,8 8 8e e e

nT T T

r r rI t I I

D t D t D t (3.12)

Page 94: Time-Dependent Mechanics of Living Cells

94

where ( ) ( )/ ( )nI t I t I is the normalized fluorescence intensity in the bleached spot

as a function of time t , er is the effective bleach radius, 0I and 1I are modified Bessel

functions. In the experimentally acquired FRAP curves, the first post-bleach value

( 1 )nI t s was larger than zero (Figure ‎5-4 and Figure ‎5-5B, C), something that could be

due to rapid fluorophore diffusion during the bleach process or incomplete weak

bleaching [141]. In cells exposed to hyperosmotic conditions, the intensity of the first

post-bleach time point was over 30% lower than in isoosmotic conditions, suggesting that

the non-zero value of this first time point was due to rapid diffusive recovery. Based on

this observation, we assigned a timing of 0t 0.1 s to the first post bleach time point

when fitting fluorescence recovery curves. The validity of the fitting was verified by

comparing the translational diffusion constant of GFP and CMFDA estimated with our

technique to previously published values.

3.11.3 Estimation of cortical F-actin turnover half-time

To assess the turnover rate of the F-actin cytoskeleton, HeLa cells stably expressing

actin-GFP were blocked in prometaphase by overnight treatment with 100 nM

nocodazole. Under these conditions, cells form a well-defined actin cortex. To estimate F-

actin turnover, photobleaching was performed on the F-actin cortex and recovery was

imaged at a frame rate of 0.9s/frame (Figure ‎5-4B).

3.12 Data processing, curve fitting and statistical analysis

Indentation and stress relaxation curves collected by AFM were analysed using custom

written code in Matlab (Mathworks Inc, Cambridge, UK). Data points where the

indentation depth d was larger than 25% of cell height were excluded. Goodness of fit

was evaluated by calculating 2r values and for analyses only fits with 2r > 0.85 were

considered (representing more than 90% of the collected data). The calculated value for

Page 95: Time-Dependent Mechanics of Living Cells

95

each group of variables is presented in terms of mean and standard deviation (Mean±SD).

To test pairwise differences of population experiments, the student’s t-test was performed

between individual treatments. Values of p < 0.01 compared to control were considered

significant and are indicated by asterisks in the graphs.

Page 96: Time-Dependent Mechanics of Living Cells

96

Chapter 4

The cytoplasm of living cells

behaves as a poroelastic material

Equation Chapter (Next) Section 1

‎The cytoplasm of living cells is constituted of a porous elastic solid ‎meshwork

(cytoskeleton, organelles, ‎macromolecules, etc) bathing in an interstitial fluid (cytosol)

(see Figure ‎1-1 and Figure ‎1-10A). ‎Based on this description, it is not very surprising that

the cytoplasm has many hydrogel-like characteristics‎. Hydrogels, an aggregate of cross-

linked polymer network and water molecules, are ubiquitous in nature and especially in

biological tissues. One of the main characteristics of hydrogels is that their mechanical

response is manifested through concurrent interaction of the solid network and the fluid

phase. In such fluid-filled sponges, the rate of relaxation is limited by the rate at which

interstitial fluid ‎can redistribute within the solid network. As briefly described in Chapter

2.3, the Biot theory of poroelasticity has been applied to study the rheology of fluid-

infiltrated porous elastic solids. In this chapter, two different experimental approaches are

utilized to examine the poroelastic behaviours of cytoplasm and compare it to hydrogels.

Page 97: Time-Dependent Mechanics of Living Cells

97

First, by employing AFM micro indentation tests acquired on cells and hydrogels, the

functional form of force-relaxation curves at different ‎timescales was examined. Then the

idea of poroelastic cytoplasm is validated by comparing the force relaxation curves

acquired on cells with those acquired on hydrogels as these are well-characterized

poroelastic ‎materials. ‎Second, to further test the poroelastic behaviour of cytoplasm, the

rheology of cells in response to different types of osmotic perturbation were studied and

the swelling/shrinking kinetics of cytoplasm were investigated in the framework of

poroelasticity.

4.1 laitanadfa‎atea‎hfe‎nfnennatedenadfa‎fh‎ feftIneadn‎inatednI‎

4.1.1 Establishing the experimental conditions to probe poroelasticity

First, I established the experimental conditions under which water redistribution within

the cytoplasm (or hydrogel) might contribute to force-relaxation. In our experiments

(Figure ‎3-2), following rapid indentation with an AFM cantilever (3.5-6nN applied during

a rise time rt ~ 35ms resulting in typical cellular indentation of d ~ 1µm), force

decreased by ~ 35% whereas indentation depth only increased by less than ~ 5%,

therefore our experiments measure force-relaxation under approximately constant applied

strain (Figure ‎4-1A). Relaxation in poroelastic materials is due to water movement out of

the porous matrix in the compressed region (Figure ‎2-7B, Figure ‎3-2C). The time-scale

for water movement is 2~ /p pt L D (L is the length-scale associated with

indentation[120]: ~L Rd with R the radius of the indenter) and therefore poroelastic

relaxation contributes significantly if the rate of force application is faster than the rate of

water efflux: r pt t . Previous experiments estimated pD ~ 1-100 µm2.s

-1 in cells [100,

101] yielding a characteristic poroelastic time of pt ~ 0.1-10 s, far longer than rt . Hence,

if intracellular water redistribution is important for cell rheology, force-relaxation curves

should display characteristic poroelastic signatures for times up to pt ~ 0.1-10 s.

Page 98: Time-Dependent Mechanics of Living Cells

98

Figure ‎4-1 Stress relaxation of HeLa cells in response to AFM microindentation

(A) Temporal evolution of the indentation depth (black) and the measured force (grey) in response to AFM

microindentation normalized to their values when the target force is reached. Inset: approach phase from

which the elasticity is calculated (grey curve). The total approach time lasts less than 50 ms. (B) cell height

distribution at the indentation point. (C) Indentation depth distribution. In B and C, N indicates the total

number of measurements and n indicates the number of cells.

For hydrogels similar to the ones that were used in our experiments, poroelastic diffusion

constants of pD ~ 100 µm2.s

-1 have been reported [137]. Considering this estimate,

applying indention depths of d ~ 2 ‎µm with the same size indentors as for cells (radius of

7.5 ‎µm) would give characteristic poroelastic times pt < 0.1 s, too fast to accurately

observe poroelastic effects, given a rise times of rt ~ 0.1 s. Therefore to be able to

reliably observe poroelastic effects, I used a larger indentor ( radius of R = 25 µm).

Using this larger indentor with d ~ 2 ‎µm gave an expected characteristic poroelastic time

of pt > 0.5 s much larger than the experimental rise time rt .

Page 99: Time-Dependent Mechanics of Living Cells

99

4.1.2 Poroelasticity is the dominant mechanism of force-relaxation in hydrogels

To test our rescaling strategies on hydrogels, force-indentation and force relaxation

curves were acquired on hydrogels by setting target forces of 40, 80 and 140 nN as shown

in Figure ‎4-2. For each set of target forces, ‎20 replicate measurements were performed on

a ‎100 μm × 100 μm area of the gel and the relaxation curves were averaged for each time

point. Analysis of force indentation curves for each set of target forces respectively yields

indentation depths of 1.99±0.02 μm, 2.77‎±0.03 μm and 3.86±0.06 μm and not

significantly different elastic moduli of 1.8±0.05 kPa, 1.9±0.06 kPa and 2±0.1 kPa.

Figure ‎4-2B shows the plot of force relaxation curves in log-log coordinates which

displays two clear plateaus at very short and at very long time scales for each set of

curves. The long timescale plateau indicates that the force relaxation in the gel has a clear

characteristic relaxation time. Furthermore these curves show that force relaxation in

these hydrogels does not exhibit a power law behaviour.

One hallmark of poroelastic materials is that their characteristic relaxation time

~ /p pt R Dd is dependent on the indentation depth d . In contrast, power law relaxations

of the form 0 0( ) ~ ( / )F t F t t b and exponential relaxations of the form

0( ) ~ exp( / )F t F t t have parameters b and t that are independent of length scale.

For ideal stress-relaxations of any power-law or linear viscoelastic material,

normalisation of force between any two arbitrarily chosen times 1t and 2t

1 2 1[ ( ) ( )]/[ ( ) ( )]F t F t F t F t should lead to collapse of all experimental relaxation

curves onto one master curve. However, for poroelastic materials, force normalisation

alone is not sufficient to collapse curves onto a single master curve and time must also be

renormalised with respect to the indentation depth to achieve collapse of curves.

Page 100: Time-Dependent Mechanics of Living Cells

100

Figure ‎4-2 Force-relaxation curves acquired with different indentation depths on hydrogel.

‎(A) Force-‎relaxation curves acquired on a hydrogel for three different indentation depths ( 1d , 2d and 3d ). Each

data point is the average of 20 ‎measurements acquired at different positions on the hydrogel surface. (B) log-

log plot of the curves shown in (A). Curves show characteristic ‎plateaus at short and long time-scales. In all

graphs, solid lines and grey shadings represent the mean and standard deviations at each data point,

respectively.

To distinguish between poroelastic and viscoelastic force-relaxations, I undertook a

rescaling analysis as explained in Chapter 3.8. First the force was normalised according to

equation (3.7) up to the time when each curve reaches its final plateau ( *t t ). As

shown in Figure ‎4-3A, normalisation of force alone did not result in the collapse of

curves onto a single master curve. If the gel were a linear viscoelastic material,

normalisation of force in this way should result in collapse of all curves onto a single

curve irrespective of indentation depth. However, if in addition to force, time is also

normalised by indentation depth, all experimental curves collapse onto a single curve,

indicating that poroelasticity is the dominant mechanism of force relaxation in hydrogels

(Figure ‎4-3B).

Page 101: Time-Dependent Mechanics of Living Cells

101

Figure ‎4-3 Normalisation of force-relaxation curves acquired on hydrogel.

(A) Normalisation of force-‎relaxation curves (Figure ‎4-2) using equation (3.7) plotted as a function of time.

For force renormalisation, a separate time *t was chosen for each curve such that *t t , with t the ‎time

at which each curve reaches its plateau: 1 *t = 7.5 s, 2 *t = 10 s and 3 *t = 12.5 s. (B) Normalised force-

relaxation curves from A plotted as a ‎function of time divided by indentation depth d ( /t d ). Time rescaling

was performed for each force relaxation curve separately using their ‎respective indentation depth. Following

force renormalisation and time-rescaling, all force-relaxation curves collapsed onto one single curve. (C)

Normalisation of force-relaxation curves shown in Figure ‎4-2 using equation (3.7) plotted as a function

of ‎time. For force renormalisation, a separate time *t dependent on indentation depth was chosen for each

curve with *t ad . Choosing a = 1.25 s.µm-‎‎1 gave times 1 *t = 2.5 s, 2 *t = 3.5 s and 3 *t = 4.9 s. (D)

Normalised force-relaxation curves from C plotted as a function of time divided by ‎indentation depth d (

/t d ). Time rescaling was performed for each force-relaxation curve separately using their respective

indentation depth. ‎Following force renormalisation and time-rescaling, all force-relaxation curves collapsed

onto one single curve. ‎In all graphs, solid lines and grey shadings represent the mean and standard deviations

at each data point, ‎respectively.

Page 102: Time-Dependent Mechanics of Living Cells

102

As explained earlier in Chapter 3.8, collapse of curves can also be obtained following

normalisation of time for times shorter than t . For times *t t ‎‎to obtain collapse of

force relaxation of poroelastic material onto a master curve, *t must be selected

proportional to the indentation depth d specific to each curve: *t ad where a (with

unit of s.μm-1

) is an arbitrary number fixed for all curves (Figure ‎4-3C). With this choice

of *t a similar collapse of curves was observed using force and time normalisation

(Figure ‎4-3D), confirming the dominance of poroelasticity in the force relaxation of gels

in these experiments. Fitting the relaxation curves with equation (3.5) gave a poroelastic

diffusion constant of 33±4 μm.s-1

for the hydrogel comparable to the previously reported

values [137].

4.1.3 Cellular force-relaxation at short timescales is poroelastic

Population averaged force-relaxation curves showed similar trends for both HeLa and

MDCK cells with a rapid decay in the first 0.5 s followed by slower decay afterwards

(Figure ‎4-4A). When plotted in log-log scale in Figure ‎4-4B, we see that force-relaxation

clearly displayed two separate regimes: a plateau lasting ~ 0.1-0.2 s followed by a

transition to a linear regime. This indicated that at short time-scales cellular force-

relaxation did not follow a simple power law. Comparison with force-relaxation curves

acquired on physical hydrogels, which display a plateau at short time-scales followed by

a transition to a second plateau at longer time-scales (Figure ‎4-2A, B), suggests that the

initial plateau observed in cellular force-relaxation may correspond to poroelastic

behaviour. Indeed poroelastic models fitted the force-relaxation data well for short times

(< 0.5 s); whereas power law models were applicable for times longer than ~ 0.1-0.2 s

(Figure ‎4-4B).

Page 103: Time-Dependent Mechanics of Living Cells

103

Figure ‎4-4 Force-relaxation curves acquired with different indentation depths on cells.

‎(A) Population averaged force-relaxation curves for HeLa cells (green) and MDCK cells (blue) for ‎target

indentation depths of 1.45 μm for HeLa cells and 1.75 μm for MDCK cells. Curves are averages of ‎n = 5

HeLa cells and n = 20 MDCK cells. The grey shaded area around the average relaxation curves ‎represents

the standard deviation of the data. (B) Population averaged force-relaxation curves for HeLa ‎cells (green) and

MDCK cells (blue) from D-I plotted in a log-log scale. For both cell types, experimental ‎force-relaxation was

fitted with poroelastic (black solid line) and power law relaxations (grey solid line).‎

To gain further insight into the nature of cellular force-relaxation at short time-scales, I

acquired experimental relaxation curves following indentations with increasing depths.

For analysis of these experiments, I selected cells with identical elasticities such that

application of a chosen target force resulted in identical indentation depths and I averaged

their relaxation curves. This filtering procedure ensured that the cellular relaxation curves

were comparable in all aspects (Figure ‎4-5A and Figure ‎4-6A) and allowed us to assess

whether or not cellular relaxation was dependent upon indentation depth. As shown

earlier for hydrogels, the most important property of these gels as a poroelastic material is

that their characteristic relaxation time depends on indentation depth. Indeed, the length

scale dependency of relaxation in poroelastic materials is in contrast to power law and

linear viscoelastic exponential relaxation behaviours.

Page 104: Time-Dependent Mechanics of Living Cells

104

Figure ‎4-5 Normalisation of force-relaxation curves acquired on MDCK cells.

(A) Force-relaxation curves acquired on MDCK cells for three different indentation depths ( 1d , 2d and 3d ).

Each relaxation curve is the average of 20 force-relaxation curves acquired on cells with matching elasticities

of E = 0.4±0.1 kPa. Elasticity matching ensured that identical forces were applied onto each cell to obtain a

given indentation depth. (B) Normalisation of force-relaxation curves shown in A using equation (3.7)

choosing *t = 3.6 s for all curves. Relaxation curves were significantly different from one another for times

shorter than ~ 0.5 s but appeared to collapse at times longer than ~ 1.2 s. (C) Normalisation of force-

relaxation curves shown in A using equation (3.7) plotted as a function of time. For force renormalisation, a

separate time *t dependent on indentation depth was chosen for each curve with *t ad . Choosing a =

0.16 s.µm-1 gave times 1 *t ~ 105 ms, 2 *t ~ 170 ms and 3 *t ~ 285 ms. (D) Normalised force-relaxation

curves from C plotted as a function of time divided by indentation depth d ( /t d ). Time rescaling was

performed for each force relaxation curve separately using their respective indentation depth d . Following

force renormalisation and time-rescaling, all three force-relaxation curves collapsed onto one single curve. In

all graphs, solid lines and grey shadings represent the mean and standard deviations at each data point,

respectively.

Page 105: Time-Dependent Mechanics of Living Cells

105

Figure ‎4-6 Normalisation of force-relaxation curves acquired on HeLa cells.

(A) Force-‎relaxation curves acquired on HeLa cells for two different indentation depths ( 1d and 2d ). Each

relaxation curve is the average of 5 force-‎relaxation curves acquired on cells with matching elasticities of E

= 0.66±0.1 kPa and heights of h = 5.5±0.3 μm. This ensured that identical forces ‎were applied onto each cell

to obtain a given indentation depth. Matching of cell heights was necessary to minimise substrate effects due

to the ‎smaller thickness of HeLa cells compared to MDCK cells. (B) Normalisation of force-relaxation curves

shown in A using equation (3.7) ‎choosing *t = 1 s for all curves. Relaxation curves were significantly

different from one another for times shorter than ~ 0.5 s but appeared to ‎collapse at times longer than ~ 0.6 s.

(C) Normalisation of force-relaxation curves shown in A using equation (3.7) plotted as a function of

time. ‎For force renormalisation, a separate time *t dependent on indentation depth was chosen for each

curve with *t ad . Choosing a = 0.19 s.µm-1 ‎gave times 1 *t ~ 170 ms and 2 *t ~ 275 ms. (D)

Normalised force-relaxation curves from C plotted as a function of time divided by indentation ‎depth d (

/t d ). Time rescaling was performed for each force relaxation curve separately using their respective

indentation depth. Following force ‎renormalisation and time-rescaling, both force-relaxation curves collapsed

onto one single curve. ‎In all graphs, solid lines and grey shadings represent the mean and standard deviations

at each data point, respectively.

Page 106: Time-Dependent Mechanics of Living Cells

106

After force normalisation, experimental force-relaxation curves acquired on cells for

different indentation depths collapsed for times longer than ~ 1 s, but were significantly

different from one another for shorter times (Figure ‎4-5B for MDCK cells and

Figure ‎4-6B for HeLa cells), suggesting that, at short times, relaxation is length-scale

dependent consistent with a poroelastic behaviour.

For cells, experimental force-relaxation curves do not reach a plateau because relaxation

follows a power law at long time-scales (Figure ‎4-4B). Hence, to determine if relaxation

displayed a poroelastic behaviour at short time-scales, I followed the force normalisation

and time rescaling steps described in Chapter 3.8 for cases where *t t . These

procedures resulted in all cellular force-relaxation curves collapsing onto a single master

curve (Figure ‎4-5D for MDCK cells and Figure ‎4-6D for HeLa cells). This behaviour was

apparent in poroelastic hydrogels for both long and short time-scales (Figure ‎4-3B, D)

and in cells for times shorter than ~ 0.5 s (Figure ‎4-5D and Figure ‎4-6D). Taken together,

these data suggested that at short time-scales cellular relaxation is indentation depth

dependent and therefore that cells behaved as poroelastic materials.

4.2 Kinetics of whole cell swelling/shrinking

Using fluorescent particles attached to the cell membrane (see Figure ‎4-7A) and particle

tracking techniques, my aim here was to monitor the response of cells to osmotic

perturbations to investigate the swelling/shrinking kinetics of the cytoplasm. In particular

off-plane defocused images of fluorescent particles (depicted in Figure ‎3-6A, C) enabled

us to measure time-dependent cellular deformations in the z-direction very precisely.

HeLa cells with fluorescent beads bound to their membrane were suddenly exposed to a

change of osmolarity induced by addition of a small volume of water/sucrose (achieving

final concentrations of 50% water and 400 mM sucrose). A sudden change in osmolarity

induces osmotic pressure gradients that drive water in or out of the cell resulting in it

swelling or shrinking. After application of hypo/hyperosmotic shock, changes in the cell

Page 107: Time-Dependent Mechanics of Living Cells

107

height were measured over time by monitoring changes in the radius of the outer ring

(formed by the out of focus image of the fluorescent bead attached to the cell surface,

Figure ‎3-6A, B, C) to determine changes in z position of the bead.

Figure ‎4-7 Experimental setup to investigate the poroelastic behaviour of cells during swelling/shrinkage.

(A-I) Schematic of fluorescent microbeads attached to the cell membrane. (A-II) By changing the osmolarity

of the extracellular medium we can induce osmotic pressure gradients across the cell membrane that drive

water in or out of the cell resulting in swelling or shrinkage (Here a decrease in cell volume due to

hyperosmotic shock is depicted). Keeping the focal plane fixed, changes in the vertical position of each bead

due to swelling/shrinkage of the cell can be determined from the outer ring radius of the defocused image

using the calibration curve, Figure ‎3-6C. (B) Changes in height of beads over time after application of hypo-

osmotic/hyperosmotic shock at time 0t . Fitting the displacement curve to the analytical solutions for

poroelastic hydrogels allows the poroelastic properties of the cell to be estimated.

First, I asked whether the cell behaves as a poroelastic gel by fitting the experimental

displacement curves to the analytical solutions for swelling kinetics of poroelastic gels.

Then the poroelastic diffusion coefficient pD was computed and compared with the

values obtained from AFM indentation experiments in Chapter 5. As proposed in Chapter

6, this method could be employed to dissect the contribution of different cellular layers to

cell rheology and to determine the spatial distribution of pD .

Page 108: Time-Dependent Mechanics of Living Cells

108

4.2.1 Swelling/shrinkage of a constrained poroelastic material

Consider a thin layer of poroelastic material (a gel) attached on one side to a fixed

substrate and immersed in a solvent for sufficiently long to reach an equilibrium

homogeneous state with 0C and 0m the initial concentration and chemical potential of

the solvent respectively. When the gel is suddenly exposed to a solvent with different

chemical potential m , the chemical potential gradient 0m m drives the solvent to flow

into/out of the gel. The displacement and chemical potential fields can be considered one

dimensional in z if the lateral dimensions xL and yL of the gel are much larger than its

thickness H ( ,x yL L H ). Indeed, in this case, there is a negligible amount of flow

from the edges and the solvent flows mostly through top surface of the gel. The gel is

constrained in xy so 0xx yye e and the problem is similar to uniaxial strain

consolidation as described in Chapter 2 (2-4-2). For cells, constraint in xy is achieved

through focal contacts and integrin transmembrane proteins. The gel can freely swell and

shrink from the top surface ( 0zzs ) and thus the equations (2.48) and (2.50) can be

written in terms of chemical potential:

= 0

0

(1 2 )( )

2 (1 )(1 2 )

( ),(1 )

szzz

s s

sxx yy

s

ua

z G

b

n m me

nn m m

s sn

(4.1)

2

2.pDt x

m m (4.2)

Considering the top surface as the reference point, the spatial and temporal evolution of

chemical potential can be derived using appropriate boundary and initial conditions. The

equilibrium chemical potential before application of osmotic shock yields the initial

condition 0( , 0)zm m . A step change in chemical potential on the top surface which

remains constant over time (0, )tm m and the no flow condition at the bottom of the

Page 109: Time-Dependent Mechanics of Living Cells

109

gel | 0z Hzm

yield the boundary conditions. Therefore the solution of the diffusion

equation (4.2) is very similar to equation (2.52):

2 2

0 1,3,..

( , ) 4sin exp ,

2n

z t n zn

m Hm m p

p tm m p

(4.3)

where 2/4pD t Ht is the dimensionless time. One can also obtain the solution for

infinite thickness (also applicable for early times after shock) as:

0

( , ).

4 p

z t zerfc

D tm m

m m (4.4)

After a long enough time 2 /t H D , the solvent inside the gel has reached a new

equilibrium m m and the gel has shrunken/swollen from its initial height H , by h to

its final height. Considering the total final strain at very long times =zz

hH

e and

employing equation (4.1)-a, the final change in thickness of the gel is:

0(1 2 )

.2 (1 )

s

s s

h HG

n m mn

(4.5)

The change of thickness at top surface ( )h t is derived by integrating equation (4.1) and

using equations (4.3) and (4.5):

2 22 2

1,3,..

( ) 81 exp

n

h tn

h mp t

p (4.6)

Using equation (4.4)‎‎the change in the thickness of the gel is written as:

( ) 2

.pD th t

h H p (4.7)

Page 110: Time-Dependent Mechanics of Living Cells

110

Figure ‎4-8 Kinetics of cell swelling/shrinking in HeLa cells

(A, B) Swelling/shrinking kinetics of HeLa cells in response to hyperosmotic (A) and hypoosmotic (B)

shocks. Dynamics of cellular swelling/shrinking is precisely probed by tracking the height ( )h t of fluorescent

beads attached on the cell membrane at different heights H . The curves show the change cell height

normalised to the final change in thickness ( )/h t h . The inset shows the absolute change of height ( )h t . (C)

Swelling/shrinking curves were fitted to the poroelastic model using equation (4.6) derived for poroelastic

swelling of constrained gels. Fits to the equation (4.7) describes the early stage behaviour of cellular

swelling/shrinking. ‎

Page 111: Time-Dependent Mechanics of Living Cells

111

4.2.2 Poroelasticity can explain the kinetics of cellular swelling/shrinking

During the experiments, HeLa cells were exposed to hyper/hypoosmotic medium at t =0

and ‎the vertical displacement of the attached beads was measured as a function of time to

obtain swelling/shrinking curves (Figure ‎4-8A, B insets). These curves were then

normalized with respect to the final change in the height h (Figure ‎4-8A,B). The

( )/h t h curves were fitted with the analytical solution derived for a poroelastic

cytoplasm in equation (4.6) (Figure ‎4-8C). This solution fitted the experimental curves

well and fitting gave the poroelastic diffusion constants of pD =‎‎0.3±0.2 ‎μm2.s

-1 (n = 16)

and pD =‎‎0.6±0.3‎‎μm2.s

-1 (‎n ‎= 10)‎‎for swelling and shrinking respectively. Furthermore,

considering the early stage behaviour, the first 10 s of curves were fitted with analytical

solution in equation (4.7) and diffusion coefficients of pD = 0.4±0.2 μm2.s

-1 and pD ‎‎=

0.7±0.3 μm2.s

-1 were obtained for swelling and shrinking respectively (Figure ‎4-8C). The

values obtained from fits with the short timescale solution were in good agreement with

those obtained fitting with the full analytical solution.

4.3 Discussion and conclusions

To validate the experimental strategy and analytical methods, first the force-relaxation

behaviour of hydrogels, which ‎are well-characterised poroelastic materials, was

examined. Intrinsic viscoelasticity and poroelasticity concurrently contribute to the time

dependent deformation of gels. The intrinsic viscoelasticity is associated with

conformational changes in the solid meshwork, while poroelasticity is related to the

redistribution of fluid. During indentation tests on gels (similar to the hydrogel that is

used in my experiments) the viscoelastic relaxation time results from the occurrence of

multiple microscopic processes such as slippage and reptation of polymer ‎chains [87].

This relaxation time is determined by the ratio of the shear viscosity h and the shear

modulus G : ~ /v Gt h

[87], and is indentation depth independent ‎as both h and G are

Page 112: Time-Dependent Mechanics of Living Cells

112

independent of contact size. In contrast, the poroelastic relaxation time ~ /p pR Dt d

is

related to the time scale of fluid redistribution/migration within/out of the deformed

region and this depends on the contact radius and indentation depth. For hydrogels, plots

of force-relaxation curves in log-log coordinates suggest the involvement of a limited

number of relaxation processes consistent with poroelastic behaviour. More importantly,

for experiments performed with different indentation depths, ‎appropriate normalisation of

experimental force-relaxation curves resulted in the collapse of all curves onto a single

master curve. This provides strong evidence for the dominance of poroelasticity in

relaxation of hydrogels in agreement with [121, 137, 142]. It may thus be concluded that

poroelasticity is the dominant mechanism of relaxation in our ‎indentation ‎experiments

on ‎hydrogels.

A detailed examination of the functional form of cellular force-relaxation curves at

different time scales was undertaken and these curves were compared to force-relaxation

curves acquired on hydrogels. Force-relaxation induced by fast localised indentation

by AFM contained two regimes: at short time-scales, the time-scale of force- relaxation

was indentation depth dependent, a hallmark of poroelastic materials; while at

longer timescales relaxation exhibited a power law behaviour. This confirms that living

cells are much more complex active materials than simple hydrogels. Indeed,

at ‎physiological ‎time-scales (0.1-10s), in addition to conformational changes in polymeric

network ‎of ‎cytoskeleton and rearrangements in the solid phase ‎(macromolecules) other

factors such as ‎turnover ‎of cytoskeletal fibres, activity of myosin molecular motors and

network ‎rearrangements ‎due to ‎association and dissociation of crosslinkers, ‎also influence

the relaxation significantly.

Cell mechanical studies over the years have revealed a rich phenomenological landscape

of rheological behaviours that are dependent upon probe geometry, loading protocol and

loading frequency [56, 58, 63-65], though the biological origin of many of these regimes

Page 113: Time-Dependent Mechanics of Living Cells

113

remains to be fully explored. It is widely recognized that over wide ranges of frequency,

the rheology of living cells exhibits a weak power law behaviour in which the viscoelastic

response does not exhibit any characteristic timescale (see Chapter 2.2.3)‎. The structural

damping model, which has previously been applied to characterise a variety of biological

materials [143], was recently implemented to describe the observed power law behaviour

in cells [33]. In the power law structural damping model, ‎the storage and loss moduli are

both weak power law functions of frequency: ' '', ~G G bw , 0 1b , and are related

through a structural damping coefficient '' '/ tan( /2)damp G Gh bp . As a result, in

this model, dissipation is linked to an elastic stress rather than an independent viscous

stress. A Newtonian viscous term m was added to the loss modulus of the model to

capture the observed dynamics at short timescales [112]. In magnetic twisting cytometry,

the addition of such a linear term could be to take into account the effects of drag forces

generated from viscous interaction between the probe and the surrounding fluid.

However, the calculated values for this viscous term ( 1m pa.s, [33, 79, 144]) are much

larger than the viscosity of surrounding fluid (~ 0.001 pa.s). Therefore, the physical

origins of this additional viscous parameter are still not clear and are therefore debatable.

The shape of my force-relaxation curves at short timescales was not consistent with a

power law behaviour (similar to experiments reported by others [53, 145-149]‎) and it is

surprising that the initial non-power law regime observed in my experiments should not

have been described before. I envisage two major reasons for this. First, most recent

experiments aim to describe the shear rheology of cells over many decades of frequency

and therefore apply oscillatory deformations to the cell, using shear rheology methods

rather than focusing on volumetric changes (dilatational rheology) as done here. Indeed,

measurements of shear rheology are of little or ‎no consequence in understanding cell

rheology for which volumetric deformations are critical (e.g. cell migration [150] and the

formation of quasi-spherical protrusions known as blebs [107]). My experiments were

designed to apply large volumetric deformations via micro-indentation tests such that

Page 114: Time-Dependent Mechanics of Living Cells

114

they inherently capture the dilatational rheology of the cell. Indeed, my experiments are

placed in a regime where ‎intracellular water flows should play an important role: rapid

application of a large volumetric deformation and ‎observation of force relaxation over

short and intermediate time-scales using microindentation. Second, in experiments

reported by other groups, such as force-‎relaxation experiments similar to the ones

presented here [52, 53, 145, 151] and also using other very different experimental setups

[146, 147, 149], a rapid initial force relaxation was observed but its origin was not

examined. In these experiments, either the short timescale behaviours were neglected,

only the asymptotic ‎part of relaxation was fitted, or functions such as double exponentials

with both fast and slow relaxation times or stretched exponential functions were used to

fit the experimental data.

To observe water flows in response to applied deformations, one needs to deform the cell

with a large probe over a short rise time rt . Indeed, the time scale of application of

deformation rt must be shorter than the time it takes for water to drain out of the

deformed volume 2 2 2~ / ~ /( )p pt L D L Em x (L is the probe length-scale and in the case

of indentation: ~L Rd with R the radius of the indenter and d indentation depth). In

the context of oscillatory sinusoidal loading, shorter rise time deformations are achieved

by increasing the frequency of excitation. However, if the loading frequency is too high,

water does not have time to drain out of the deformed volume before the next cycle of

deformation is applied. In this regime, water will not move relative to the mesh and

therefore inertial and structural viscous effects from the mesh are sufficient to describe

the system [81, 152]. Previous work has shown that for frequencies 2 2/( )pf E Lx m

such coupling occurs [81]. Hence, to observe poroelastic effects in oscillatory

experiments, it is required to deform the material at frequencies larger than 1/ pt to

satisfy the short rise time condition to induce fluid flow but smaller than pf so that fluid

and mesh are not coupled. As both conditions have identical scaling, it is expected that

they only be satisfied in a narrow window around pf ~ 5 Hz for my experimental

Page 115: Time-Dependent Mechanics of Living Cells

115

conditions. Most oscillatory loading experiments sample rheology in 2 points per decade

and hence would likely not detect this window. Hence, to observe poroelastic effects, we

need to apply deformations with rise times shorter than pt and repeated at frequencies

lower than pf .

It should be emphasised that this type of loading regime is very physiologically relevant.

Indeed, cells in tissues of the cardiovascular or respiratory systems are exposed to large

strain deformations (tens of %) applied at high strain rates (> 20%.s-1

) but with low repeat

frequencies (< 4 Hz) (e.g. 10% strain applied at ~ 50%.s-1

repeated at up to 4 Hz for

arterial walls [153], 70% strain applied at up to 900%.s-1

repeated at up to 4 Hz in the

myocardial wall [154], and 20% strain applied at > 20%.s-1

repeated at ~ 1 Hz for lung

alveola [155]). As the changes in cell shape that occur in physiological tissue deformation

are similar in magnitude or larger than those induced in our experiments and are repeated

with low frequency, I expect intracellular water redistribution to play an important role in

the relaxation of cells within these tissues

When force-relaxation curves acquired for different indentation depths on cells were

renormalized for force and rescaled with a time-scale dependent on indentation depth, all

experimental curves collapsed onto a single master curve for short time-scales,

confirming that the initial dynamics of cellular relaxation is dominated by poroelastic

effects and intracellular water flow (Figure ‎4-5 and Figure ‎4-6), but for larger timescales

the relaxation curves exhibit a weak power law behaviour. Together these data suggest

that in my experiments the power-law dynamics and poroelasticity coexist. For timescales

shorter than ~ 0.5 s, intracellular water redistribution contributed strongly to force-

relaxation, consistent with the ~ 0.1 s time-scale measured for intracellular water flows in

the cytoplasm of HeLa cells [156]. At long timescales when fluid is fully redistributed,

fluid related poroelastic dissipations became negligible and structural damping effects

(manifested in the form of a weak power law) come into play as the main source of

dissipation.

Page 116: Time-Dependent Mechanics of Living Cells

116

Volume exclusion effects are the primary source of swelling/shrinking ‎behaviour in

artificial and biological gels as well as living cells. ‎The dynamics of swelling and

shrinking mainly depends on the rate at which water can be transported ‎into and out of

the gel. However, poroelasticity can also significantly affect the swelling/shrinking

kinetics: for an elastic porous medium, the deformation of the porous matrix induced by

osmotic shock significantly limits the rate of swelling/shrinkage. Indeed a very stiff

sponge cannot swell/shrink even if water can diffuse into/out of it infinitely fast. Here the

hypothesis that the swelling/shrinking behaviour of cytoplasm is similar to that of

hydrogels was tested by examining its volume response to osmotic perturbations. It is

shown that that the swelling/shrinking kinetics of cytoplasm can be well described via

a ‎simple 1-D poroelastic model and analytical solutions fitted the experimental data well.

The estimated values of the diffusion coefficient were within the previously reported

range of 0.1 to 10 μm.s-1‎ [100-102] but around one order of magnitude smaller than those

obtained from indentation experiments (as presented in next chapter). There are several

possible sources of this discrepancy as shall be explained in the following paragraph.

A one dimensional model for swelling/shrinking of a constrained gel was used to

determine the cellular poroelastic diffusion coefficient from fits of experimental curves.

This one dimensional formulation was a good approximation for my analysis because the

lateral ‎movements of the membrane-attached beads were much smaller than their vertical

displacement (less than ~ 0.5 μm in xy versus greater than ~ 2 μm in z ). ‎However this

model ‎is only valid if the osmotic perturbations are applied in a step manner (infinitely

fast)‎. In our ‎experiments, it takes a certain amount of time for the added osmolyte to mix

with the bath solution and for osmolyte to equilibrate. This non instantaneous change in

osmolarity could be one reason for underestimation of diffusion ‎coefficient using

proposed the poroelastic model. Comparing shrinking and swelling experiments, ‎faster

poroelastic diffusion constants were obtained in shrinking experiments, which might be

due to ‎irreversible dynamics of the membrane water channels (aquaporins) or different

Page 117: Time-Dependent Mechanics of Living Cells

117

responses of the solid ‎phase and cytoskeleton to tensional ‎versus compressional forces

created during hyperosmotic ‎shrinkage and hypoosmotic swelling, respectively. ‎

The structural heterogeneity and complex nature of the cell may be the key reason that

different experimental setups report different values for cellular mechanical and

rheological properties. AFM probes only small regions of cytoplasm whereas all of the

structures inside the whole cell are perturbed during a swelling/shrinking experiment.

Indeed several cellular layers‎, such as the cell membrane, the cell cortex, the nucleus, and

the granular filamentous interior, each with different permeabilities and elasticities,

contribute to setting the poroelastic behaviour of the cell in swelling/shrinking. During

indentation experiments, small regions of cytoplasm underneath the indentor are

compressed and‎ water movement is directed towards the cell interior (see Figure ‎4-9). For

these indentation experiments, the cell membrane most likely does not play a significant

role in setting the rate of water permeation and the membrane can be assumed to be fully

permeable. In contrast, during swelling/shrinkage of the whole cell, water infiltration

through cytoplasm and all parts of the cell (including the nucleus) sets the dynamics of

swelling/shrinking. Specifically the rate of swelling/shrinking can be limited significantly

by the rate at which water can pass through thin layers of cytoplasm such as the cell

cortex and the plasma membrane.

Page 118: Time-Dependent Mechanics of Living Cells

118

Figure ‎4-9 Schematic representation of intracellular water movements in different experimental setups.

During osmotic shocks the water permeates through whole cytoplasm and all parts of the cell whereas for

AFM indentation tests water movements is restricted to small regions of cytoplasm underneath the indentor.

The arrows show possible directions of water movement.

The plasma membrane has long been thought of as the main regulator for volumetric

responses of cells to osmotic perturbations. However, it has been recently shown that in

the absence of a cell membrane the cytoplasm of ‎mammalian cells has intrinsic

osmosensitivity [157]. Simple diffusion through the lipid bilayer and water selective

facilitated diffusion through the water channels (aquaporins) are the main pathways for

water permeation during osmotic perturbations [156, 158]. Coherent anti-Stokes Raman

scattering (CARS) microscopy allows observation of intracellular hydrodynamics by

monitoring H2O/D2O exchange in living cells [159]. Using CARS, recent experimental

measurements found that for Hela S3 cells it takes less than ~ 100 ms for water to fully

permeate across the plasma membrane and fully exchange, much shorter than our

measured relaxation times (>> 1 s) for the volumetric response of HeLa cells to osmotic

perturbations. Therefore, in addition to diffusional effects, mechanical factors also play

an important role in setting the relaxation time. Furthermore, considering a cell height of

~ 4 μm and a poroelastic diffusion coefficient ~ 40 μm2.s

-1 yields the ratio of /pD h ‎‎~ 10

μm.s-1

which is of the magnitude as the measured membrane diffusional permeability ‎~

10-40 μm.s-1

[156, 159] suggesting that the solid phase of cytoplasm has a similar

Page 119: Time-Dependent Mechanics of Living Cells

119

contribution in limiting the rate of water permeation to the membrane. I conclude that the

simple concept of a water-filled container surrounded by membrane, cannot fully explain

the volumetric response of cell to osmotic perturbations and that the solid phase of

cytoplasm contributes significantly in this response. I suggest that poroelastic model for

the cytoplasm is a more physical concept that inherently considers the gel-like nature of

cytoplasm and can help explain the dynamics of volume changes. However, further

experiments suggested in Chapter 6 are required to precisely dissect the role of the

membrane and different parts of the solid phase in determining the cell’s response to

osmotic shocks.

Page 120: Time-Dependent Mechanics of Living Cells

120

Chapter 5

Poroelastic properties of

cytoplasm: Osmotic perturbations

and the role of cytoskeleton ‎

Equation Chapter (Next) Section 1

It has been shown in the previous chapter that the framework of poroelasticity is well

suited to understanding the time-dependent mechanical properties of cells. In this

framework, the cytoplasm comprises two phases: a porous solid phase (cytoskeleton,

organelles, macromolecules) and a fluid phase (cytosol). The conceptual advantage of

such a description is that cellular rheology can be related to measurable cellular

parameters (elastic modulus E , hydraulic pore size x , and cytosolic viscosity m ) using a

simple scaling law 2~ /pD Ex m and therefore changes in rheology resulting from

changes in cellular organisation can be qualitatively predicted. In this chapter, AFM

indentation tests were utilized to measure cellular stress-relaxation in response to rapid

Page 121: Time-Dependent Mechanics of Living Cells

121

local force application in conjunction with chemical, genetic, and osmotic treatments to

modulate two of the parameters that scaling law predicts should influence cellular time-

dependent mechanical properties: pore size x and the elasticity E .

5.1 Determination of cellular elastic, viscoelastic and poroelastic

parameters from AFM measurements

To provide baseline behaviour for perturbation experiments, I measured the elastic,

viscoelastic and poroelastic properties of MDCK, HeLa, and HT1080 cells by fitting

force-indentation (Figure ‎3-3) and force-relaxation (Figure ‎5-1) curves with Hertzian and

viscoelastic or poroelastic models, respectively. Measurement of average cell thickness

suggested that for time-scales shorter than 0.5 s, cells could be considered semi-infinite

and forces relaxed according to an exponential relationship with 2( ) exp( / )pF t D t L

(see Chapter 2.4, 5, Figure ‎2-5 and Figure ‎2-7C). In our experimental conditions (3.5-6

nN force resulting in indentation depths less than 25% of cell height, Figure ‎4-1B, C),

force-relaxation with an average amplitude of 40% was observed with 80% of total

relaxation occurring in ~ 0.5 s (Figure ‎4-4 and Figure ‎5-1A).

Analysis of the indentation curves yielded an elastic modulus of E = 0.9±0.4 kPa for

HeLa cells (N = 189 curves on n = 25 cells), E = 0.4±0.2 kPa for HT1080 cells (N =

161 curves on n =27 cells), and E = 0.4±0.1 kPa for MDCK cells (n = 20 cells).

Poroelastic models fitted experimental force-relaxation curves well (on average 2r =

0.95, black line, Figure ‎5-1A) and yielded a poroelastic diffusion constant of pD = 41±11

µm2.s

-1 for HeLa cells, pD = 40±10 µm

2.s

-1 for HT1080 cells, and pD = 61±10 µm

2.s

-1 for

MDCK cells.

Page 122: Time-Dependent Mechanics of Living Cells

122

Figure ‎5-1 Fitting force-relaxation curves.

(A) The first 0.5 s of force-relaxation curves were fitted with the poroelastic model (black line) and the

standard linear viscoelastic model (grey line). Poroelastic models fitted the data better than the single phase

viscoelastic model especially at short times. Inset: the percentage error of each fit defined as | |

. (B) Ratio of root mean square error (RMSE) of the viscoelastic fit to the RMSE of poroelastic fit.

Assuming that the cytoplasm behaves as a standard linear viscoelastic material and to

measure the apparent viscosity of the cytoplasm, I fitted force-relaxation curves with

equation (3.6) ( 1( ) ~ exp( / )F t k t h ). Using the approach phase of AFM indentation

curves, I measured cellular elasticity and assumed 1 ~k E . Therefore the relaxations

predicted by a viscoelastic formulation depend on only one free parameter: the apparent

viscosity h . Since the force-relaxation curves are indentation depth dependent, to

calculate the apparent viscosity I considered only the curves with indentation depths of

±10% of the mean values obtained from distribution of indentation depths (Figure 4-1B).

Viscoelastic formulations were found to replicate experimental force-relaxation curves

well (on average 2r = 0.92), but the early phase of force-relaxation was fit somewhat less

accurately by viscoelastic models (gray line, Figure ‎5-1A) compared to poroelastic fits

(black line, Figure ‎5-1A). Indeed, viscoelastic models gave rise to somewhat larger errors

Page 123: Time-Dependent Mechanics of Living Cells

123

than poroelastic models (Figure ‎5-1B). An average apparent viscosity of h = 166±81 Pa.s

was measured for HeLa cells, ‎h ‎= 77±37 Pa.s for HT1080 cells and h = 50±15 Pa.s for

MDCK cells.

5.2 Poroelasticity can predict changes in stress relaxation in

response to volume changes

Having established that poroelastic models provide good fits for stress relaxation in cells,

I examined the ability of our simple scaling law to qualitatively predict changes in pD

resulting from changes in pore size due to cell volume change. Indeed, cell volume

changes should not affect cytoskeletal organisation or integrity but should alter

cytoplasmic pore size. In our experiments, we exposed Hela and MDCK cells to media of

different osmolarities: hyperosmotic media to decrease cell volume and hypoosmotic

media to increase cell volume.

First, I measured cell volume change in response to changes in osmolarity for up to 30

min to ascertain that volume changes persisted long enough for experimental

measurements to be carried out with AFM (Figure ‎5-2A). Indeed, cell volume stayed

constant over this time frame. I verified that photobleaching due to the acquisition of

confocal stacks over an extended period of time did not artefactually affect our

measurement of cell volume (control, n = 6 cells, Figure ‎5-2A). Cells regulate their

volume tightly to be able to tolerate extracellular environments of different osmolarities.

When osmolarity is perturbed, flow of water across cell membrane causes cell swelling or

shrinkage. However, the cell can restore its original volume by activation of channels or

transporters (typically K+, Cl and organic osmolytes) to lose or gain sufficient amount of

osmolytes to return to its preferred volume. To ensure a stable volume increase under

hypoosmotic conditions over the timescale necessary for AFM experiments (~ 30 min),

cells were treated simultaneously with regulatory volume decrease (RVD) inhibitors

Page 124: Time-Dependent Mechanics of Living Cells

124

[160]. Also, MDCK cells were treated with regulatory volume increase (RVI)

inhibitors ‎to ensure a stable decrease in cell volume after shrinkage. Under these

conditions, a stable increase of 22±2% in cell volume was measured after hypoosmotic

treatment (n = 3 cells, Figure ‎5-2A). Conversely, upon addition of 240 mM sucrose, cell

volume decreased by 21±6% (n = 10 cells, Figure ‎5-2A) and upon addition of PEG-400

(30% volumetric concentration), cell volume decreased drastically to 46±3% of the

original volume (N = 6 cells, Figure ‎5-2A)†.

Next, using AFM indentations I investigated whether changes in cell volume resulted in

changes in the poroelastic diffusion constant. Increases in cell volume resulted in a

significant increase (55%) in poroelastic diffusion constant pD and in a significant

decrease (-20%) in cellular elasticity E (N = 92 measurements on n = 33 cells, p <

0.01 when compared to control for both E and pD , Figure ‎5-2B). In contrast, cells

exposed to PEG-400 exhibited a two-fold lower diffusion constant ( p < 0.01 when

compared to controls) than control cells and a 50% larger elasticity ( p < 0.01, N = 176

measurements on n = 17 cells) (Figure ‎5-2B). Addition of sucrose with low

concentrations (<<250 mM) resulted in no significant changes in elasticity or diffusion

constant ( p = 0.53 for E and p = 0.89 for pD ).

As will be examined in detail in next section, there are spatial inhomogeneities in shape

architecture and mechanical properties of HeLa cells spread on glass substrates. To

increase the accuracy of the measurements and decrease the uncertainties arising from

spatial inhomogeneities, similar experiments were performed on MDCK cells with 8 data

points corresponding to sampling different medium osmolarities (Figure ‎5-2C).

Qualitatively, both cell types had the same behaviour, confirming the generality of our

results (Figure ‎5-2C, D).

† I would like to acknowledge the preliminary data from report of L. Valon.

Page 125: Time-Dependent Mechanics of Living Cells

125

Figure ‎5-2 Poroelastic and elastic properties change in response to changes in cell volume.

In all graphs, error bars indicate the standard deviation and, in graphs B and C, asterisks indicate significant

changes compared to control ( p < 0.01). N indicates the total number of measurements and n indicates the

number of cells. In hypoosmotic shock experiments, cells were incubated with NPPB and DCPIB (N+D on

the graph), inhibitors of RVD. In hyperosmotic shock experiments, MDCK cells were incubated with

RVI inhibitors (EIPA). (A) Cell volume change over time in response to changes in extracellular osmolarity.

The volume was normalized to the initial cell volume at t = 0 s. The arrow indicates the time of addition of

osmolytes. (B) Effect of osmotic treatments on the elasticity E (squares) and poroelastic diffusion constant

pD (circles) in HeLa cells. (C) Effect of osmotic treatments on the elasticity E (squares) and poroelastic

diffusion constant pD (circles) in MDCK cells. (D) pD and E plotted as a function of the normalised

volumetric pore size 1/30~ ( / )V Vy in log-log plots for MDCK cells (black squares and circles) and HeLa

cells (grey squares and circles). Straight lines were fitted to the experimental data points weighted by the

number of measurements to reveal the scaling of pD and E with changes in volumetric pore size (grey

lines for HeLa cells, 1.6~E y and 2.9~pD y and black lines for MDCK cells, 5.9~E y and 1.9~pD y ).

Page 126: Time-Dependent Mechanics of Living Cells

126

Consistent with results reported by others [125], our experiments revealed that cells

relaxed less rapidly and became stiffer with decreasing fluid fraction. Increases in cell

volume resulted in a significant increase in poroelastic diffusion constant pD and a

significant decrease in cellular elasticity E (Figure ‎5-2B, C). In contrast, a decrease in

cell volume decreased the diffusion constant and increased elasticity (Figure ‎5-2B, C).

Because the exact relationship between hydraulic pore size x and cell volume is

unknown, pD and E were plotted as a function of the change in the volumetric pore size

1/30~ ( / )V Vy . As shown in the log-log plot in Figure ‎5-2D, for both MDCK and HeLa

cells, pD scaled with y and the cellular elasticity E scaled inversely with y . In

summary, an increase in cell volume increased the poroelastic diffusion constant and

decrease in cell volume decreased pD qualitatively consistent with our simple scaling

law.

To exclude any possible indirect effect of volume change due to changes in cytoskeletal

organisation, I verified that cytoskeletal structure was not perturbed by changes in cell

volume (Figure ‎5-3 for actin).

Page 127: Time-Dependent Mechanics of Living Cells

127

Figure ‎5-3 Effects of hypoosmotic or hyperosmotic shock on HeLa cells F-actin structural organization.

The F-actin cytoskeleton does not reorganise in response to hypoosmotic treatment (A) or hyperosmotic

treatment (B). All images are maximum projections of a confocal image stack of cells expressing Life-act

ruby (red). Nuclei were stained with Hoechst 34332 (blue). In A and B, image (I) shows the cell before

addition of osmolyte and image (II) after addition. Scale bar = 10 µm. No dramatic changes in structure of

actin filaments could be observed after hypoosmotic (A) or hyperosmotic (B) shock.

5.3 Changes in cell volume result in changes in cytoplasmic pore

size.

Having shown that changes in cell volume result in changes in the poroelastic diffusion

constant without affecting cytoskeletal structure, I tried to directly detect if the volume

perturbations resulted in changes of pore size. As shall be described in the following,

experiments examining the mobility of microinjected quantum dots (qdots) and

photobleaching experiments were performed to show that osmotic perturbations and the

concomitant volume changes directly affect the cytoplasmic pore size. Furthermore, as

poroelasticity is related to fluid redistribution phenomena, I verified the existence of a

Page 128: Time-Dependent Mechanics of Living Cells

128

fluid phase in cells under extreme shrinkage conditions by photobleaching of small

cytoplasmic fluorescein analogs.

5.3.1 Hyperosmotic shock halts movement of quantum dots inside cytoplasm

PEG-passivated qdots (nanoparticles made of a semiconductor material with unique

optical and electrical properties, ~ 14 nm hydrodynamic radius with the PEG-passivation

layer [161]) were microinjected into cells and their mobility was examined before and

after application of hyperosmotic shock. Under isoosmotic condition, qdots rapidly

diffused throughout the cell; however, upon addition of PEG-400, they became immobile

(supplementary video 1, Figure ‎5-5A-I, n = 7 cells examined). Hence, cytoplasmic pore

size decreased in response to cell volume decrease trapping qdots in the cytoplasmic solid

fraction and immobilising them (Figure ‎5-5A-II). This suggested that the isoosmotic pore

radius x was larger than 14 nm, consistent with our estimates from poroelasticity (see

section 5.5).

Page 129: Time-Dependent Mechanics of Living Cells

129

Figure ‎5-4 Fluorescence recovery after photobleaching in HeLa cells.

‎(A) ‎Fluorescence recovery after photobleaching of monomeric free diffusing GFP in isosmotic conditions.

Solid lines ‎indicate fluorescence recovery after photobleaching and are the average of ‎ N ‎= 18 measurements

on ‎n ‎= 7 cells. The experimental recovery was fitted with the theoretical model described in Chapter 3.11

(grey line). ‎‎(B) Fluorescence recovery after photobleaching of GFP-actin in the cortex of HeLa cells blocked

in ‎prometaphase. Solid lines indicate fluorescence recovery after photobleaching and are the average of‎ ‎N =

10 measurements on ‎n = 4 cells.‎ A half time recovery of 1/2, F actint = 11.2±3 s was obtained.‎In (A) and (B),

dashed lines indicate loss of ‎fluorescence due to imaging in a region outside of ‎the zone of FRAP, error bars

indicate the standard ‎deviation for each time point, and the greyed ‎area indicates the duration of

photobleaching. ‎

5.3.2 Translational diffusion of fluorescent probes affected by osmotic

perturbations

5.3.2.1 Effective bleach radius in FRAP experiments.

To derive quantitative measurements of the translational diffusion constants of molecules

within the cytoplasm using the methods described in [141], I experimentally determined

the effective bleach radius in the first post-bleach image for a nominal bleaching zone of

diameter 0.5 µm (Figure ‎3-7III). Averaging over all diameters passing through the centre

Page 130: Time-Dependent Mechanics of Living Cells

130

of the nominal bleach regions yielded an experimental curve that could be fit with

equation (3.11) to find er = 1.5±0.2 µm (S5A-V, N = 6 measurements on n = 6 cells).

5.3.2.2 Measurement of translational diffusion constants in cells.

To validate my estimations of the translational diffusion constant TD using the analytical

procedures described in Chapter 3.11, I first measured TD for GFP molecules and small

fluorescein analogs (CMFDA, hydrodynamic radius hR ~ 0.9 nm [162]) freely diffusing

in the cytoplasm. For cytoplasmic GFP molecules, I found ,T GFPD ~ 25 µm2.s

-1

(Figure ‎5-4A, N = 18 measurements on n = 7 cells), consistent with values reported by

others [141]. I also estimated a translational diffusion constant ,T CMFDAD ~ 38 µm2.s

-1 for

CMFDA (N = 19 measurements on n = 7 cells) which is in a good agreement with

published values in cells (Figure ‎5-5B, TD = 24-40 µm2.s

-1, [162]). Indeed, effects of

molecular crowding and hindered diffusion inside the cell slow the diffusion of GFP and

CMFDA by more than ~3 fold compared to diffusion of these analogs in aqueous

solutions ( , ,T GFP aqueousD ~ 90 µm2.s

-1 and , ,T CMFDA aqueousD ~ 240 µm

2.s

-1 [162, 163]). It

should be noted that with CMFDA loss of fluorescence due to imaging is very high (also

reported by others, Figure ‎5-5B). Therefore, only data points up to a total fluorescence

loss of 15% due to imaging were utilised to estimate the translational diffusion of

fluorophores in cells in normal isotonic condition.

5.3.2.3 Osmotic perturbations change the cytoplasmic pore size

In isoosmotic conditions, CMFDA recovered rapidly after photobleaching (black line,

Figure ‎5-5B, Table 5-1) whereas recovery slowed three fold in the presence of PEG-400

consistent with [164] (grey line, Figure ‎5-5B and Table ‎5-1). Recovery after

photobleaching of CMFDA under strong hyperosmotic conditions confirmed the presence

of a fluid-phase indicating that under severe hyperosmotic conditions cells still retained a

significant fluid fraction. To estimate the reduction in translational diffusion of CMFDA

Page 131: Time-Dependent Mechanics of Living Cells

131

in cells under hyperosmotic conditions, I normalised the data following the procedures

described in [165] and this gave an estimated value of , ,T CMFDA HyperD ~ 14 µm2.s

-1 (N =

20 measurements on n = 5 cells). The measured decrease in translational diffusion

suggested a reduction in the cytoplasmic pore size with dehydration. Indeed, translational

diffusion is related to the solid fraction f via the empirical relation / ~ exp( )T TD D f

with TD the translational diffusion constant of the molecule in a dilute isotropic

solution [166]. Assuming that the fluid is contained in N pores of equal radius x , the

solid fraction is 3~ /( )s sV V Nf x with sV the volume of the solid fraction, a constant.

/T TD D is therefore a monotonic increasing function of x . Therefore the decrease in

TD measured in response to hyperosmotic shock suggests a decrease in x . To examine

the effect of volume increase on pore size, I examined the fluorescence recovery after

photobleaching of a cytoplasmic GFP decamer (EGFP-10x, hR ~ 7.5 nm). Increases in

cell volume resulted in a significant ~ 2-fold increase in TD ( p < 0.01, Figure ‎5-5C,

Table 5-1), suggesting that pore size did increase. Together, our experiments show that

changes in cell volume modulate cytoplasmic pore size x consistent with estimates from

AFM measurements (Figure ‎5-11).

Table ‎5-1 Effects of osmotic perturbations on translational diffusion coefficients.

Translation diffusional coefficients TD are reported with unit of (µm2.s-1).

GFP

Control

N = 18, n = 7

CMFDA

Control

N = 19, n = 7

CMFDA

PEG-400

N = 20, n = 5

EGFP10x

Control

N = 17, n = 6

EGFP10x

Water+N+D

N = 23, n = 7

24.9‎±‎7 38.3±8 13.7±4 8.6±2 15.5±5

Page 132: Time-Dependent Mechanics of Living Cells

132

Figure ‎5-5 Changes in cell volume change cytoplasmic pore size.

(A) Movement of PEG-passivated quantum dots microinjected into HeLa cells in isoosmotic conditions (I)

and in hyperosmotic conditions (II). Both images are a projection of 120 frames totalling 18 s (Supplementary

movie). In isoosmotic conditions, quantum dots moved freely and the time-projection appeared blurry (I);

whereas in hyperosmotic conditions, quantum dots were immobile and the time-projected image allowed

individual quantum dots to be identified (II). Images A-I and II are single confocal sections. In (B) and (C),

dashed lines indicate loss of fluorescence due to imaging in a region outside of the zone where fluorescence

recovery after photobleaching (FRAP) was measured, solid lines indicate fluorescence recovery after

photobleaching and are the average of N measurements and error bars indicate the standard deviation for

each time point. The greyed area indicates the duration of photobleaching. (B) FRAP of CMFDA (a

fluorescein analog) in isoosmotic (black, N = 19 measurements on n = 7 cells) and hyperosmotic

conditions (grey, N = 20 measurements on n = 5 cells). In both conditions, fluorescence recovered after

photobleaching but the rate of recovery was decreased significantly in hyperosmotic conditions. (C) FRAP of

EGFP-10x (a GFP decamer) in isoosmotic (black, N = 17 measurements on n = 6 cells) and hypoosmotic

conditions (grey, N = 23 measurements on n = 7 cells). The rate of recovery was increased significantly in

hypoosmotic conditions.

Page 133: Time-Dependent Mechanics of Living Cells

133

5.4 Estimation of the hydraulic pore size from experimental

measurements of Dp.

To estimate HeLa cells hydraulic pore size, equation (2.38), 2 /p sD Ea x m , can be

used. As a first approximation, I considered sn ~ 0.3 and a fluid fraction of j ~ 0.75 for

cells in isotonic conditions which yields a Kozney constant of k ~ 4 assuming the lower

bound for a random distribution of particulate spheres [167]. These values yield an

estimate of ~ 0.05 for a in isotonic conditions. Multiple experimental reports have

shown that for small molecules the viscosity m (~ 0.001 Pa.s) of the fluid-phase of

cytoplasm (cytosol) is 2-3 fold higher in cells than in aqueous media [71, 168]. Using

2 /p sD Ea x m , together with our experimental measurements of pD ~ 40 µm2.s

-1 and

sE ~ 1 kPa in HeLa cells yields a hydraulic pore size x of ~ 15 nm, comparable to the

value estimated in section 5.3.

5.5 Spatial variations in poroelastic properties

Maps of cellular elasticity measured by AFM show that E is strongly dependent upon

location within the cell with actin-rich organelles (lamellipodium, actin stress fibres)

appearing significantly stiffer than other parts of the cell [169]. Hence, I asked if

something similar could be observed for poroelastic properties by measuring the cellular

poroelastic properties at different locations along the cell long axis (Figure ‎3-5). I

performed N = 386 total measurements on n = 30 cells and displayed the measurements

averaged over 2 µm bins as a function of distance to the centroid of the nucleus. Cell

height decreased significantly with increasing distance from the nucleus (Figure ‎5-6A)

and this measurement enabled us to exclude low areas of the cell that are prone to

measurement artefacts due to the limited thickness. The poroelastic diffusion constant pD

decreased slightly, but not significantly, away from the nucleus going from 40 µm2.s

-1 to

Page 134: Time-Dependent Mechanics of Living Cells

134

~ 30 µm2.s

-1 ( p > 0.02, Figure ‎5-6C). In contrast, elasticity increased significantly away

from the nucleus increasing from ~ 500 Pa to ~ 1200 Pa in lower areas (Figure ‎5-6B).

Finally, the lumped pore size estimated from the ratio 1/2/pD Em also decreased

significantly away from the nucleus (Figure ‎5-6D).

Figure ‎5-6 Spatial distribution of cellular rheological properties.

(A-D) Scatter plots of the average cell height h , elasticity E , poroelastic diffusion constant pD , and

lumped pore size 1/2

/pD Em as a function of distance to the centre of nucleus. Each scatter graph is

generated by averaging values of each variable over 2 μm wide bins. Filled black circles indicate locations

above the nucleus, white unfilled circles in the cytoplasm, and grey filled circles boundary areas. The lines

indicate the weighted least-square fit of the scattered data. Unfilled circles with a cross superimposed indicate

measurements significantly different ( p < 0.01) from those obtained above the nucleus.

Page 135: Time-Dependent Mechanics of Living Cells

135

5.6 Poroelastic properties are influenced by cytoskeletal integrity

It has been shown previously that cellular poroelastic properties are affected by the cell

cycle state of the cell [101]. Specifically, in mitotic cells blocked in metaphase ( F-actin

and vimentin intermediate filaments concentrate at the cell periphery, microtubules

reorganize to form the mitotic spindle and cytokeratin filaments disassemble) a ~ two fold

larger poroelastic diffusion constant was measured compared to interphase cells [101]. It

has been hypothesized that this may be due to the dramatic reorganization of the

cytoskeleton concomitant with metaphase [101]. Cytoskeletal organisation strongly

influences cellular elasticity E [169-171], and is also likely to affect the cytoplasmic

pore size x (Figure ‎5-7A-C). As both factors (elasticity and pore size) play antagonistic

roles in setting pD , in the following I examined the effect of disruption of the

cytoskeleton on the cellular poroelastic properties using chemical and genetic treatments.

5.6.1 Chemical treatments

Treatment of cells with 750 nM latrunculin, a drug that depolymerises the actin

cytoskeleton, resulted in a significant decrease in the cellular elastic modulus

(Figure ‎5-9A, consistent with [169]), a ~ two-fold increase in the poroelastic diffusion

coefficient (Figure ‎5-9B), and a significant increase in the lumped pore size

(Figure ‎5-11C). Depolymerisation of microtubules by treatment with 5 µM nocodazole

had no significant effect (Figure ‎5-9 and Figure ‎5-11C). Stabilisation of microtubules

with 350 nM taxol resulted in a small (-16%) but significant decrease in elasticity but did

not alter pD (Figure ‎5-9 and Figure ‎5-11C). Perturbation of contractility with the myosin

II ATPase inhibitor blebbistatin (100 μM) lead to a 50% increase in pD , a 70% decrease

in E , and a significant increase in lumped pore size (Figure ‎5-9 and Figure ‎5-11C)‡.

‡ I would like to acknowledge the preliminary data from report of L. Valon.

Page 136: Time-Dependent Mechanics of Living Cells

136

Figure ‎5-7 Cellular distribution of cytoskeletal fibres.

Scale bars = 10 µm. Nuclei are shown in blue and GFP-tagged cytoskeletal fibres are shown in green. (A)

Maximum projection of a cell expressing tubulin-GFP. GFP-α-tubulin was homogenously localised

throughout the cell body. (B) Maximum projection of cells expressing Life-act-GFP, an F-actin reporter

construct. F-actin was enriched in cell protrusions but was also present in the cell body. (C) Maximum

projection of cells expressing keratin18-GFP. Intermediate filaments (keratin 18) were homogenously

distributed throughout the cell body. (D) Overexpression of the dominant mutant Keratin 14 R125C resulted

in aggregation of the cellular keratin network. (A&B were acquired by L. Valon).

5.6.2 Genetic modification

In light of the dramatic effect of F-actin depolymerisation on pD and x , we attempted to

decrease the pore size by expressing a constitutively active mutant of WASp (WASp

I294T, CA-WASp) that results in excessive polymerization of F-actin in the cytoplasm

through ectopic activation of the arp2/3 complex, an F-actin nucleator (Figure ‎5-8, [26]).

Increased cytoplasmic F-actin due to CA-WASp resulted in a significant decrease in the

Page 137: Time-Dependent Mechanics of Living Cells

137

poroelastic diffusion coefficient (-43%, p < 0.01, Figure ‎5-9B), a significant increase in

cellular elasticity (+33%, p < 0.01, Figure ‎5-9A), and a significant decrease in the

lumped pore size (-38%, p < 0.01, Figure ‎5-11C). Similar results were also obtained for

HT1080 cells (-49% for pD , +68% for E and -38% for lumped pore size, p < 0.01).

Figure ‎5-8 Ectopic polymerization of F-actin due to CA-WASp.

HeLa cells were transduced with a lentivirus encoding GFP-CA-WASp (in green) and stained for F-actin with

Rhodamine-Phalloidin (in red). Cells expressing high levels of CA-WASp (green, I) had more cytoplasmic F-

actin (red, II, green arrow) than cells expressing no CA-WASp (II, white arrow). (III) zx profile of the cells

shown in (I) and (II) taken along the dashed line. Cells expressing CA-WASp displayed more intense

cytoplasmic F-actin staining (green arrow) than control cells (white arrow). Cortical actin fluorescence levels

appeared unchanged. Nuclei are shown in blue. Images I and II are single confocal sections ‎acquired with

kind helps from Dr. Moulding. Scale bars = 10 µm.

Page 138: Time-Dependent Mechanics of Living Cells

138

Figure ‎5-9 Effect of drug treatments and genetic perturbations on poroelastic properties.

Effect of F-actin depolymerisation (Latrunculin treatment), F-actin overpolymerisation (overexpression of

CA-WASp), myosin inhibition (blebbistatin treatment), microtubule stabilisation (Taxol treatment),

microtubule over-polymerisation (overexpression of γ-tubulin), microtubule depolymerisation (nocodazole

treatment), intermediate filaments aggregation (expression of Keratin 14 R125C) and crosslinking

perturbation (overexpression of ΔABD-α-actinin) on the cellular elasticity E in (A) and poroelastic

diffusion constant pD in (B). Asterisks indicate significant changes ( p < 0.01). N is the total number of

measurements and n the number of cells examined.

Page 139: Time-Dependent Mechanics of Living Cells

139

Then, we attempted to change cell rheology without affecting intracellular F-actin

concentration by perturbing crosslinking or contractility. To perturb crosslinking, we

overexpressed a deletion mutant of α-actinin (ΔABD-α-actinin) that lacks an actin-

binding domain but can still dimerize with endogenous protein [172], reasoning that this

should either increase the F-actin gel entanglement length l or reduce the average

diameter of F-actin bundles b (the inverse scenario is depicted in Figure ‎5-12B).

Overexpression of ΔABD-α-actinin led to a significant decrease in E but no change in

pD or lumped pore size (Figure ‎5-9 and Figure ‎5-11C). To determine if ectopic

polymerisation of microtubules had similar effects to CA-WASp, we overexpressed γ-

tubulin, a microtubule nucleator [173], but found that this had no effect on cellular

elasticity, poroelastic diffusion constant, or lumped pore size (Figure ‎5-9 and

Figure ‎5-11C). Finally, expression of a dominant keratin mutant (Keratin 14 R125C,

[130]) that causes aggregation of the cellular keratin intermediate filament network had

no effect on E , pD , or lumped pore size (Figure ‎5-9 and Figure ‎5-11C).

5.6.2.1 ‎Clinical relevance: Mechanical disruption of mitosis in cells with

dysregulated F-actin ‎production:

Dynamic events during cell division, such as the poleward movement of chromosomes or

the ‎closure of the cleavage furrow, are mechanical processes that are intrinsically

linked ‎to the physical properties of the cell. It is well established that actin is the main

cytoskeletal determinant of cellular ‎rheology (as in this thesis under framework of

poroelasticity), and provides the physical support to maintain ‎cell shape and drive shape

change [169-171]. In some patients, malfunctioning of the WASp protein can lead to

severe immunodeficiency, known as x-linked neutropenia. CA-WASp (a constitutive

WASp activation) causes ‎increased F-actin production through dysregulated activation of

the Arp2/3 complex throughout the cytoplasm, which results in loss of proliferation, high

levels of apoptosis, cell division defects and chromosomal instability [26, 27]. In a very

recent study, we investigated the role of WASp in cell division and asked if an increased

Page 140: Time-Dependent Mechanics of Living Cells

140

abundance of F-actin in CA-WASp could cause these defects (Moulding et al, Blood,

2012‎). We showed that excessive cytoplasmic F-actin increases the elasticity and

apparent viscosity of the cytoplasm and mechanically impedes chromosome separation

and furrow closure during mitosis. Inhibition of the Arp2/3 complex reduced the change

in viscosity, restored the velocity of chromosome movement and reversed the XLN

phenotype. Taken together, these results show that non-specific alterations to cellular

rheology can lead to disease by perturbing critical functions such as mitosis.

Figure ‎5-10 Cell rheology and disease.

AFM measurements on HT1080 cells revealed a ~ two-fold increase in both cellular elasticity and apparent

cellular viscosity due to CA-WASp expression. Inhibition of the Arp2/3 complex using different

concentration of the CK666 drug reduced the amount of cytoplasmic F-actin in CA-WASp cells, leading to

partial recovery of the cell rheological properties and reducing cell division defects (Moulding et al, Blood,

2012)

Page 141: Time-Dependent Mechanics of Living Cells

141

5.7 Discussion and conclusions

I have shown that water redistribution plays a significant role in cellular responses to

mechanical stresses and that the effect of osmotic and cytoskeletal perturbations on

cellular rheology can be understood in the framework of poroelasticity through a simple

scaling law 2~ /pD Ex m . I tested the dependence of pD on the hydraulic pore size x by

modulating cell volume and showed that pD scaled proportionally to volume change,

consistent with a poroelastic scaling law. Changes in cell volume did not affect

cytoskeletal organisation but did modulate pore size. Experiments monitoring the

mobility of microinjected quantum dots suggested that x was ~ 14 nm consistent with

estimates based on measured values for pD and E . I also confirmed that cellular

elasticity scaled inversely with volume change as shown experimentally [125, 174] and

theoretically [84] for F-actin gels and cells. However, the exact relationship between the

poroelastic diffusion constant pD and the hydraulic pore size x could not be tested

experimentally because the relationship between volumetric change and change in x is

unknown due to the complex nature of the solid phase of the cytoplasm (composed of the

cytoskeletal gel, organelles, and macromolecules, Figure ‎5-12A, [175]). Taken together,

the results show that, for time-scales up to ~ 0.5 s, the dynamics of cellular force-

relaxation is consistent with a poroelastic behaviour for cells and that changes in cellular

volume resulted in changes in pD due to changes in x .

Although the cytoskeleton plays a fundamental role in modulating cellular elasticity and

rheology, our studies (it should be note that the indentation method that is used here only

probes local rheological properties of the cell) show that microtubules and keratin

intermediate filaments do not play a significant role in setting cellular rheological

properties (Figure ‎5-9 and Figure ‎5-11C). In contrast, both the poroelastic diffusion

constant and elasticity strongly depended on actomyosin (Figure ‎5-9). The experiments

Page 142: Time-Dependent Mechanics of Living Cells

142

qualitatively illustrate the relative importance of x and E in determining pD .

Depolymerising the F-actin cytoskeleton decreased E and increased pore size resulting

in an overall increase in pD . Conversely, when actin was ectopically polymerised in the

cytoplasm by arp2/3 activation by CA-WASp (Figure ‎5-8), E increased and the pore size

decreased resulting in a decrease in pD . For both perturbations, changes in pore size x

dominated over changes in cellular elasticity in setting pD . For dense crosslinked F-actin

gels, theoretical relationships between the entanglement length l and the elasticity E

suggest that 2 5~ /( )BE k Tk l with k the bending rigidity of the average F-actin bundle

of diameter b , Bk the Boltzmann constant, and T the temperature [84] (Figure ‎5-12A).

If the hydraulic pore size x and the cytoskeletal entanglement length l were identical,

pD would scale as 2 3~ /( )p BD k Tk m l implying that changes in elastic modulus would

dominate over changes in pore size, in direct contradiction with our results. Hence, x and

l are different and x may be influenced both by the cytoskeleton and macromolecular

crowding [175] (Figure ‎5-12A).

To decouple changes in elasticity from gross changes in intracellular F-actin

concentration, we decreased E by reducing F-actin crosslinking through overexpression

of a mutant α-actinin [172] that can either increase the entanglement length l or decrease

the bending rigidity k of F-actin bundles by diminishing their average diameter b

(Figure ‎5-12). Overexpression of mutant α-actinin led to a decrease in E but no

detectable change in pD or x confirming that pore size dominates over elasticity in

determining cell rheology. Finally, myosin inhibition led to an increase in pD , a decrease

in E , and an increase in x , indicating that myosin contractility participates in setting

rheology through application of pre-stress to the cellular F-actin mesh [176], something

that results directly or indirectly in a reduction in pore size (Figure ‎5-12B, [176, 177]).

Taken together, these results show that F-actin plays a fundamental role in modulating

cellular rheology but further work will be necessary to understand the relationship

Page 143: Time-Dependent Mechanics of Living Cells

143

between hydraulic pore size x , cytoskeletal entanglement length l , crosslinking, and

contractility in living cells.

Figure ‎5-11 Effect of volume changes, chemical and genetic treatments on the lumped pore size.

Effect of volume changes, on lumped pore size of HeLa cells in (A) and MDCK cells in (B). (C) Effect of

drug treatments and genetic perturbations on the lumped pore size of Hela cells. The lumped pore size was

estimated from the ratio 1/2( / )pD Em . Asterisks indicate significant changes ( p < 0.01). N is the total number

of measurements and n the number of cells examined.

Page 144: Time-Dependent Mechanics of Living Cells

144

Recent experiments using magnetic bead twisting rheometry have demonstrated the

existence of at least two regimes in the rheology of living cells [80, 178] and in vitro

cytoskeletal networks of actomyosin [176]. At high frequencies (5 Hz - 10 kHz), the

cellular dynamic modulus increased with frequency; whereas at lower frequencies (0.1-

5Hz) it appeared to be frequency independent. At high frequencies (~ 1 kHz ), fast

thermal fluctuations give rise to bending of cytoskeletal filaments and determine the

cellular mechanical properties [80]. However, at longer time-scales (0.2 s to hundreds of

s) cytoskeletal turnover or water movement may influence cell rheology. Measurements

of cell rheology obtained using different techniques over time-scales of 1-100 s reveal the

existence of different regimes: one with a short time-scale (~ 1 s) and the other with a

long time-scale (~ 10 s) [51, 52, 68]. Over the time-scales of the experiments described in

this thesis (~ 5 s), other factors such as turnover of cytoskeletal fibres and cytoskeletal

network rearrangements due to crosslinker exchange or myosin contractility might also in

principle influence cell rheology. In the cells studied here, F-actin, the main cytoskeletal

determinant of cellular rheology (Figure ‎5-9 and [62, 179]), turned over with a half-time

of ~ 11 s (Figure ‎5-4B), crosslinkers turned over in ~ 20 s [180], and myosin inhibition

led to faster force-relaxation. Hence, active biological remodelling cannot account for the

dissipative effects observed in our force-relaxation experiments. Taken together, the time-

scale of force-relaxation, the functional form of force-relaxation, and the qualitative

agreement between the theoretical scaling of pD with experimental changes to E and x

support our hypothesis that water redistribution is the principal source of dissipation at

short time-scales contributing ~ 50% of total relaxation in ~ 0.2 s in our experiments.

Page 145: Time-Dependent Mechanics of Living Cells

145

In both viscoelastic and poroelastic models, force decays exponentially with time during

stress-relaxation (see equations (3.5) and (3.6)). Hence, equating the exponents and

recalling 1 ~k E , I obtain the following relationship for a length-scale dependent

effective cellular viscosity h :

2

~L

h mx

(5.1)

with L the characteristic length-scale and m the viscosity of cytosol. The dependence of

η on the ratio of a mesoscopic length-scale to a microscopic length-scale in the system

may explain the large spread in reported measurements of cytoplasmic viscosities [58, 68]

(‎ranging from ~ 0.1 to 500 pa.s using different rheological measurements such as

magnetic twisting cytometry [33] and AFM indentation [51, 54] experiments)‎.

Within the poroelastic regime, further intuition for the complexity and variety of length

scales involved in setting cellular rheology (Figure ‎5-12A) can be gained by recognising

that the effective viscosity m experienced by a particle diffusing in the cytosol will

depend on its size (Figure ‎5-5, [70, 72]). Within the cellular fluid fraction, there exists a

wide distribution of particle sizes with a lower limit on the radius set by the radius of

water molecules. Whereas measuring the poroelastic diffusion constant pD remains

challenging, the diffusivity mD of any given particle can be measured accurately in cells

with experimental techniques such as the presented FRAP technique. For a molecule of

radius a , the Stokes-Einstein relationship gives /(6 )m BD k T apm or

( ) /(6 )B ma k T D am p . Any interaction between the molecule and its environment (e.g.

reaction with other molecules, crowding and hydrodynamic interactions [181], size-

exclusion [70]) will result in a deviation of the experimentally determined mD from this

relationship. Using the relationship for elasticity of gels 2 5~ /( )BE k Tk l and recalling

that the bending rigidity of filaments scales as 4~ polymerE bk (with b the average

Page 146: Time-Dependent Mechanics of Living Cells

146

diameter of a filament –or bundle of filaments-, and polymerE the elasticity of the

polymeric material [182]), yields the relationship

22 8

2 5~

( )polymer

p mB

aE bD D

k T

xl

(5.2)

in which four different length-scales contribute to setting cellular rheology. We see that

the average filament bundle diameter b , the size of the largest particles in the cytosol a

(and indeed the particle size distribution in the cytosol), the entanglement length l , and

the hydraulic pore size b conspire to determine the geometric, transport, and rheological

complexity of the cell (Figure ‎5-12A). Since all these parameters can be dynamically

controlled by the cell, it is perhaps not surprising that a rich range of rheologies has been

experimentally observed in cells [32, 33, 56, 58, 63-65, 68, 79, 80].

Page 147: Time-Dependent Mechanics of Living Cells

147

Figure ‎5-12 Schematic representation of the cytoplasm and effects of different cellular perturbations.

(A) The cytoskeleton and macromolecular crowding participate in setting the hydraulic pore size through

which water can diffuse. The length-scales involved in setting cellular rheology are the average filament

diameter b , the size a of particles in the cytosol, the hydraulic pore size x , and the entanglement length l

of the cytoskeleton. (B) Effects of different perturbations on cell rheology. (I, III) Reduction in cell volume

(I) or increase in F-actin concentration (III) causes a decrease in cytoskeletal mesh size l and an increase in

crowding which combined lead to a decrease in hydraulic pore size x . (II, IV) Increase in activity of myosin

motors or crosslinkers lead to either contraction of F-actin network (decrease in mesh size) or increase in

bending rigidity of actin cytoskeleton, which combined results in an increase in the elasticity E of the

network and a decrease in the hydraulic pore size x .

Page 148: Time-Dependent Mechanics of Living Cells

148

Chapter 6

General discussion and future

work

6.1 Further discussion

6.1.1 Cell rheology and its complexity

Living cells are complex materials displaying a high degree of structural hierarchy and

heterogeneity coupled with active biochemical processes that constantly remodel their

internal structure. Therefore, perhaps unsurprisingly, they display an astonishing variety

of rheological behaviours depending on amplitude, frequency and spatial location of

loading. Over the years, a rich phenomenology of rheological behaviours has been

uncovered in cells such as scale-free power law rheology, strain stiffening, anomalous

diffusion, and rejuvenation (see Chapter 1.4.3 and [56, 58, 63-65]).

Several theoretical models have been proposed to explain the observed behaviours but

finding a unifying theory has been difficult because different microrheological

Page 149: Time-Dependent Mechanics of Living Cells

149

measurement techniques excite different modes of relaxation [58]. These rheological

models range from those that treat the cytoplasm as a single phase material whose

rheology is described using networks of spring and dashpots [68, 79], to the sophisticated

soft glassy rheology (SGR) models that describe cells as being akin to soft glassy

materials close to a glass transition [33, 80]; in either case, the underlying geometrical

and biophysical phenomena remain poorly defined [58, 63, 64]. Furthermore, none of the

models proposed, account for dilatational changes in the multiphase material that is the

cytoplasm; these volumetric deformations are ubiquitous in the context of phenomena

such as blebbing, cell oscillations or cell movement [100, 150], and in gels of purified

cytoskeletal proteins [32] whose macroscopic rheological properties depend on the gel

structural parameters and its interaction with an interstitial fluid [81, 84, 176]. Any

unified theoretical framework that aims to capture the rheological behaviours of cells and

links these behaviours to cellular structural and biological parameters must account for

both the shear and dilatational effects seen in cell mechanics as well as account for the

role of crowding and active processes in cells.

6.1.2 Why poroelasticity?

The flow of water plays a critical part in such processes. Recent experiments suggest that

pressure equilibrates slowly within cells (~10s) [100, 101, 103] giving rise to intracellular

flows of cytosol [150, 183] that cells may exploit to create lamellipodial protrusions [150]

or blebs [100, 102] for locomotion [184]. Furthermore, the resistance to water flow

through the soft porous structure serves to slow motion in a simple and ubiquitous way

that does not depend on the details of structural viscous dissipation in the cytoplasmic

network. Based on these observations, a coarse-grained biphasic description of the

cytoplasm as a porous elastic solid meshwork bathed in an interstitial fluid (e.g.

poroelasticity [116] or the two-fluid model [185]) has been proposed as a minimal

framework for capturing the essence of cytoplasmic rheology [59, 100, 101]. In the

framework of poroelasticity, coarse graining of the physical parameters dictating cellular

Page 150: Time-Dependent Mechanics of Living Cells

150

rheology accounts for the effects of the interstitial fluid and related volume changes,

macromolecular crowding and a cytoskeletal network [100, 101, 105], consistent with the

rheological properties of the cell on the time-scales needed for redistribution of

intracellular fluids in response to localised deformation. The response of cells to

deformation then depends only on the poroelastic diffusion constant pD , with larger

values corresponding to more rapid stress relaxations. This single parameter scales as

2~ /pD Ex m , allowing changes in cellular rheology to be predicted in response to

changes in E , x , and m .

The principal shortcoming of most models of cell rheology is that they do not relate the

measured ‎rheological properties to structural or biological parameters within the cell

meaning that they lack predictive power. As a coarse-grained bottom-up approach,

one ‎advantage of using poroelasticity over other phenomenological models is that it

allows ‎cellular rheology to be related to observable physical and biological parameters,

making it mechanistic and thus giving it predictive ‎power (see Figure 6-1). In this work, I

sought to understand cell rheology at physiologically relevant time-‎scales and to attempt

to understand the cellular biological and physical variables that affect it. ‎By nature,

cellular rheology is bound to be extremely complex but my goal was to capture

its ‎essence with as few physical parameters as possible. The ‎poroelastic model of the

cytoplasm presented here is a first step in this direction. It takes as inputs

measurable ‎geometric and physical parameters and leads to falsifiable predictions ‎(see

Figure 6-1)‎. This represents a ‎conceptual advance and a step towards describing the

experimental observations with the fewest ‎physiologically and biologically relevant

parameters possible.

Page 151: Time-Dependent Mechanics of Living Cells

151

Figure ‎6-1 Microstructural parameters and mesoscopic length scale set the poroelastic time scale.

Different timescales associated with different rheological models are involved in setting cell rheology. The

time-dependent mechanics of cells arise from the combination of several distinct rheological behaviours.

These rheological behaviours give rise to different dynamics that coexist but each dominates over a range of

timescales. At very short time scales, the behaviour of network of semiflexible filaments (entropic and

enthalpic fluctuations) sets the rheology of the cytoplasm [80]. At long timescales, the cytoplasm exhibits

inelastic structural rearrangements and behaves as a soft glassy material [80]. At very long timescales the cell

can no longer be considered as a passive material and active processes such as molecular motor activities and

cytoskeletal filament turnover come into play. Depending on the length scale and time scale, cytosolic flows

and intracellular fluid redistribution (poroelastic effects) can affect the rheological behaviour of the cell. The

mesoscopic length scale of the system L , the elastic modulus of cell E , the hydraulic pore size of the

cytoplasm x , and the cytosolic viscosity m set the characteristic poroelastic timescale pt ( ~ 0.2pt s in my

AFM indentation experiments).

Time dependent mechanics of cells is the manifestation of several cell rheological

behaviours and these rheological behaviours give rise to different dynamics that could

coexist but each dominates on different timescales. In particular, poroelastic cellular

behaviours, in which dissipations are due to water redistribution, are both timescale and

lengthscale dependent allowing us to tune the experimental setup in such a way that we

observe weak or strong poroelastic effects accordingly. The idea of designing this study

was to consider a minimal poroelastic model under optimised experimental conditions,

and determine whether this minimal model was sufficient to grasp the essence of cellular

rheology and predict changes in rheology in response to perturbations. Extra complexity

could be added to the model to take into account more complex properties of the

cytoskeleton and cytoplasm or network tension: we could consider a solid phase with

Page 152: Time-Dependent Mechanics of Living Cells

152

intrinsic viscoelasticity and/or granularity instead of a simple linear elastic meshwork.

However, this would come at the expense of introducing more parameters and extra

uncertainty as one would have to make many more assumptions about cell mechanics and

the end result would not necessarily draw a clearer description of cell rheology and its

origins. In particular, because the exact relationship between fluid fraction, volumetric

pore size , hydraulic pore size , and cytoskeletal entanglement length is still

unknown for living cells, rather than making complex hypotheses about the poroelastic

scaling law, the trends of my experimental results were validated.

Future studies should concentrate on implementing more realistic cytoskeletal and

cytoplasmic properties. However addition of extra complexity should be coupled with

experimental determination/validation of all newly introduced variables and this might

require much more sophisticated experiments.

6.1.3 Direct physiological relevance

To gain an understanding of how widespread poroelastic effects are in the rheology of

isolated cells ‎and cells within tissues, we can compute the poroelastic Péclet number

eP ( )/ pVL D , with V a ‎characteristic velocity (due to active movement, external

loading, etc). For eP large compared to 1, ‎poroelastic effects dominate the viscoelastic

response of the cytoplasm to shape change due to ‎externally applied loading or intrinsic

cellular forces. In isolated cells, poroelastic effects have been ‎implicated in the formation

of protrusions such as lamellipodia or blebs [100]. For these, the rate of protrusion growth

can be chosen as a characteristic velocity. In rapidly moving cells, forward-directed

intracellular water flows [150, 183] resulting from pressure gradients due to myosin

contraction of the cell rear have been proposed to participate in lamellipodial protrusion

[150, 186]. Assuming a representative lamellipodium length of L ~10 µm and protrusion

velocities of V ~ 0.3 µm.s-1

, poroelastic effects will play an important role if pD ≤ 3

µm2.s

-1, lower than measured in the cytoplasm but consistent with the far higher F-actin

Page 153: Time-Dependent Mechanics of Living Cells

153

density observed in electron micrographs of the lamellipodium [184]. Furthermore, cells

can also migrate using blebbing motility [107] where large quasi-spherical blebs (L

~10µm) arise at the cell front with protrusion rates of V ~ 1 µm.s-1

giving an estimate of

pD ~ 10 µm

2.s

-1 (comparable to the values reported in previous Chapter ) to obtain eP

1.

During normal physiological function, cells within tissues of the respiratory and

cardiovascular systems are subjected to strains e > 20% applied at strain rates te > 20%.

s-1

. As a first approximation, we assume that these cells, with a representative length cellL ,

undergo a length change cell~L Le applied with a characteristic velocity cell~ tV Le . For

cells within the lung alveola [155], cellL ~ 30 µm, e ~ 20%, te

~ 20%. s-1

and assuming

pD ~ 10 µm

2.s

-1 (based on our measurements), we find eP

~ 3. Hence, these simple

estimates of eP suggest that water redistribution participates in setting the rheology of

cells within tissues under normal physiological conditions.

6.2 Future work

6.2.1 Stress and pressure wave propagation in cells

It is well established that cells can sense mechanical forces, react to them, and adapt to

them [58]. However, how these forces are sensed and in particular what physical variable

is detected by cells remains unclear. A better understanding of cell rheology will allow us

to predict the intracellular stress distribution induced by extracellular application of force.

Page 154: Time-Dependent Mechanics of Living Cells

154

Figure ‎6-2 Stress and pressure wave propagation in cells

Application of step force/pressure via AFM cantilever (A) or through micropipette (B). Upon application of

force/pressure the stress and pressure propagate intracellularly. (C-I) The displacement of each fluorescent

bead in the z direction is monitored as a function of time and the bead distance from the force/pressure

application point is measured. The bead displacement relaxation time increases with distance from the source.

(C-II) Expected curve for the relaxation time as a function of distance. This curve can provide information

about how fast stress/pressure propagates through the cytoplasm. Comparing these experimental data with

viscoelastic and poroelastic models will provide information about the physical nature of force transmission

inside the cytoplasm.

In the future, matching the temporal evolution of this intracellular stress distribution with

biological readouts (such as the intracellular calcium concentration or the turnover rates

of cytoskeletal proteins) should allow us to better understand what physical variable cells

sense and how it is sensed at the molecular level prior to conversion into a biochemical

signal. In particular, recent experiments suggest that poroelastic effects lead to slow stress

propagation within living cells [103]. In an attempt to understand how stresses propagate

inside the cytoplasm, future work will try to apply localized force or pressure onto the

cell via an AFM cantilever or through a micropipette (see Figure ‎6-2A, B). The particle

Page 155: Time-Dependent Mechanics of Living Cells

155

tracking technique as described in Chapter 3.10 will be used to monitor motion of

attached beads at different distances from the point of exerted force/pressure. Studying

the relaxation curves as a function of distance (see Figure ‎6-2C) will provide information

about how fast stress/pressure propagates through the cytoplasm. Comparing these

experimental data with viscoelastic and poroelastic models will provide useful

information about which physical mechanism are effectively involved in force

transmission inside the cytoplasm.

6.2.2 Cells as multi-layered poroelastic materials

From a hydrodynamic perspective, the cell can be viewed as a multilayer ‎composite

composed of plasma membrane attached to the subcortical actin gel surrounding

a ‎molecularly crowded environment saturated with cytosol. One future challenge will be

to dissect the contribution of each layer of the cell to cell rheology. ‎In Chapter 5, AFM

indentation tests identified the actin cytoskeleton as the main component of the cytoplasm

that affected the poroelastic rheology of the cell. It would be interesting to investigate the

contribution of the different layers of the cell to the swelling/shrinking experiments

presented in Chapter 4.2. In particular, my preliminary experimental results (not

presented here) suggested that the plasma membrane can play a significant role in setting

the dynamics of cell swelling/shrinking.

I propose to use defocusing microscopy to study the kinetics of cell swelling/shrinking

under different types of perturbations to investigate the multi-layered poroelastic nature

of the cell and to see if the contribution of each layer in cell rheology can be dissected.

First, to examine the role of membrane permeability, cells can be treated with small

concentration of mild detergents or pore forming proteins such as Digitonin or

Amphotericin B to make small holes in the membrane. It should be noted that after

membrane permeabilisation the cell will start swelling. As examined in Chapter 5,

swelling can affect the poroelastic properties (specifically pore size) significantly and

therefore to counterbalance the effects of cell swelling upon membrane permeabilisation,

Page 156: Time-Dependent Mechanics of Living Cells

156

a small concentration of osmolyte such as sucrose must be added to retain the isoosmotic

cell volume. I would expect the poroelastic diffusion coefficient to increase when the

membrane is permeabilised but the extent to which it increases will tell us about the

importance of membrane. Next, the effects of macromolecular crowding can be envisaged

by using higher concentrations of detergents to create holes large enough for

macromolecules to escape (up to micrometres in size). It should be noted that creation of

large holes in membrane might result in clearance of some of the basic components of the

cytoskeleton (such as actin monomers) that will affect the structure of filaments and

consequently the the cellular poroelastic properties.

Second, I will investigate the contribution of the actin cytosleketon to swelling/shrinking

dynamics by using drugs such as latrunculin B to depolymerise the actin filaments or

performing genetic treatments (as explained in Chapter 3.2) to induce excessive

polymerization of F-actin in the cytoplasm. Treatment of cells with drugs such as

latrunculin B will affect the structure of the whole actin cytoskeleton including the

organization of cortical actin. Therefore, to study the role of cell cortex in setting cell

rheology and poroelastic properties separately and in particular to solely perturb structure

and organization of the cortical actin, it would be very interesting to conduct more

targeted perturbations such as shRNA gene depletion to knock down the protein Diaph1

(a protein shown by other members of the laboratory to be responsible for polymerisation

of actin filaments in the cortex) or using the drug CK666 to inhibit the Arp2/3 complex

that alsoregulates the structure of the actin network.

Page 157: Time-Dependent Mechanics of Living Cells

157

6.3 Conclusions

The main conclusions arising from this work may be summarised as follows:

- The time-dependent mechanics of living cells exhibit poroelastic behaviours similar

to poroelastic physical gels.

- The dynamics of cell swelling and shrinking can be well described under the

framework of poroelasticity.

- Poroelasticity predicts changes in rheology due to cell volume changes and

cytoskeletal disruption.‎

- F-actin and molecular crowding are the main cytoskeletal contributors of poroelastic

cellular rheology and together ‎they set the rheological properties of the cell.‎

- In my experiments, the observed short timescale ‎(0.1-1 s) ‎cellular viscosity is due to

water redistribution in the cell.‎

Page 158: Time-Dependent Mechanics of Living Cells

158

References

[1] Boal DH (2002) Mechanics of the Cell (Cambridge Univ Pr).

[2] Howard J (2001) Mechanics of motor proteins and the cytoskeleton.

[3] Alberts B, Johnson A, Lewis J, Raff M, Roberts K, & Walter P (2002) Molecular biology of

the cell (Garland Science).

[4] Mofrad MRK & Kamm RD (2006) Cytoskeletal Mechanics - Models and Measurements.

(Cambridge University Press).

[5] Bray D (2001) Cell movements: from molecules to motility (Routledge).

[6] Desai A & Mitchison TJ (1997) Microtubule polymerization dynamics. Annual review of cell

and developmental biology 13(1):83-117.

[7] Brangwynne CP, MacKintosh FC, Kumar S, Geisse NA, Talbot J, Mahadevan L, Parker KK,

Ingber DE, & Weitz DA (2006) Microtubules can bear enhanced compressive loads in

living cells because of lateral reinforcement. The Journal of cell biology 173(5):733.

[8] Herrmann H, Bär H, Kreplak L, Strelkov SV, & Aebi U (2007) Intermediate filaments: from

cell architecture to nanomechanics. Nature Reviews Molecular Cell Biology 8(7):562-

573.

[9] Kölsch A, Windoffer R, Würflinger T, Aach T, & Leube RE (2010) The keratin-filament

cycle of assembly and disassembly. Journal of Cell Science 123(13):2266.

[10] Goldman RD, Clement S, Khuon S, Moir R, Trejo-Skalli A, Spann T, & Yoon M (1998)

Intermediate filament cytoskeletal system: dynamic and mechanical properties. Biological

Bulletin 194(3):361-363.

[11] Theriot JA, Mitchison TJ, Tilney LG, & Portnoy DA (1992) The rate of actin-based motility

of intracellular Listeria monocytogenes equals the rate of actin polymerization. Nature

357(6375):257-260.

[12] Theriot JA & Mitchison TJ (1991) Actin microfilament dynamics in locomoting cells.

Nature 352(6331):126-131.

[13] Watanabe N & Mitchison TJ (2002) Single-molecule speckle analysis of actin filament

turnover in lamellipodia. Science 295(5557):1083.

[14] Tseng Y, Kole TP, Lee JSH, Fedorov E, Almo SC, Schafer BW, & Wirtz D (2005) How

actin crosslinking and bundling proteins cooperate to generate an enhanced cell

mechanical response. Biochemical and biophysical research communications 334(1):183-

192.

[15] Edlund M, Lotano MA, & Otey CA (2001) Dynamics of α-actinin in focal adhesions and

stress fibers visualized with α-actinin-green fluorescent protein. Cell motility and the

cytoskeleton 48(3):190-200.

[16] Sanger JM, Mittal B, Pochapin MB, & Sanger JW (1987) Stress fiber and cleavage furrow

formation in living cells microinjected with fluorescently labeled α‐actinin. Cell motility

and the cytoskeleton 7(3):209-220.

[17] Lieleg O, Claessens MMAE, & Bausch AR (2010) Structure and dynamics of cross-linked

actin networks. Soft Matter 6(2):218-225.

[18] Bendix PM, Koenderink GH, Cuvelier D, Dogic Z, Koeleman BN, Brieher WM, Field CM,

Mahadevan L, & Weitz DA (2008) A quantitative analysis of contractility in active

cytoskeletal protein networks. Biophysical journal 94(8):3126-3136.

[19] Geiger B, Spatz JP, & Bershadsky AD (2009) Environmental sensing through focal

adhesions. Nature Reviews Molecular Cell Biology 10(1):21-33.

Page 159: Time-Dependent Mechanics of Living Cells

159

[20] Chen CS (2008) Mechanotransduction–a field pulling together? Journal of Cell Science

121(20):3285-3292.

[21] Hoffman BD, Grashoff C, & Schwartz MA (2011) Dynamic molecular processes mediate

cellular mechanotransduction. Nature 475(7356):316-323.

[22] Engler AJ, Sen S, Sweeney HL, & Discher DE (2006) Matrix elasticity directs stem cell

lineage specification. Cell 126(4):677-689.

[23] Trappmann B, Gautrot JE, Connelly JT, Strange DGT, Li Y, Oyen ML, Stuart MAC, Boehm

H, Li B, & Vogel V (2012) Extracellular-matrix tethering regulates stem-cell fate. Nature

materials 11(7):642-649.

[24] Li YS, Haga JH, & Chien S (2005) Molecular basis of the effects of shear stress on vascular

endothelial cells. J Biomech 38(10):1949-1971.

[25] Malek AM, Alper SL, & Izumo S (1999) Hemodynamic shear stress and its role in

atherosclerosis. JAMA: the journal of the American Medical Association 282(21):2035.

[26] Moulding DA, Blundell MP, Spiller DG, White MRH, Cory GO, Calle Y, Kempski H,

Sinclair J, Ancliff PJ, & Kinnon C (2007) Unregulated actin polymerization by WASp

causes defects of mitosis and cytokinesis in X-linked neutropenia. The Journal of

experimental medicine 204(9):2213-2224.

[27] Thrasher AJ & Burns SO (2010) WASP: a key immunological multitasker. Nature Reviews

Immunology 10(3):182-192.

[28] Goldmann WH, Galneder R, Ludwig M, Xu W, Adamson ED, Wang N, & Ezzell RM

(1998) Differences in elasticity of vinculin-deficient F9 cells measured by magnetometry

and atomic force microscopy. Experimental cell research 239(2):235-242.

[29] Lekka M, Laidler P, Gil D, Lekki J, Stachura Z, & Hrynkiewicz A (1999) Elasticity of

normal and cancerous human bladder cells studied by scanning force microscopy.

European Biophysics Journal 28(4):312-316.

[30] Cross SE, Jin YS, Rao J, & Gimzewski JK (2007) Nanomechanical analysis of cells from

cancer patients. Nature nanotechnology 2(12):780-783.

[31] Iyer S, Gaikwad R, Subba-Rao V, Woodworth C, & Sokolov I (2009) Atomic force

microscopy detects differences in the surface brush of normal and cancerous cells. Nature

nanotechnology 4(6):389-393.

[32] Bausch AR & Kroy K (2006) A bottom-up approach to cell mechanics. Nature Physics

2(4):231-238.

[33] Fabry B, Maksym GN, Butler JP, Glogauer M, Navajas D, & Fredberg JJ (2001) Scaling the

microrheology of living cells. Physical review letters 87(14):148102.

[34] Wang N, Butler JP, & Ingber DE (1993) Mechanotransduction across the cell surface and

through the cytoskeleton. Science 260(5111):1124-1127.

[35] Bausch AR, Ziemann F, Boulbitch AA, Jacobson K, & Sackmann E (1998) Local

measurements of viscoelastic parameters of adherent cell surfaces by magnetic bead

microrheometry. Biophysical journal 75(4):2038-2049.

[36] Dai J & Sheetz MP (1995) Mechanical properties of neuronal growth cone membranes

studied by tether formation with laser optical tweezers. Biophysical journal 68(3):988-

996.

[37] Yost MJ, Simpson D, Wrona K, Ridley S, Ploehn HJ, Borg TK, & Terracio L (2000) Design

and construction of a uniaxial cell stretcher. American Journal of Physiology-Heart and

Circulatory Physiology 279(6):H3124-H3130.

[38] Ellis E, McKinney J, Willoughby K, Liang S, & Povlishock J (1995) A new model for rapid

stretch-induced injury of cells in culture: characterization of the model using astrocytes.

Journal of neurotrauma 12(3):325-339.

[39] Usami S, Chen HH, Zhao Y, Chien S, & Skalak R (1993) Design and construction of a

linear shear stress flow chamber. Annals of biomedical engineering 21(1):77-83.

[40] Lu H, Koo LY, Wang WM, Lauffenburger DA, Griffith LG, & Jensen KF (2004)

Microfluidic shear devices for quantitative analysis of cell adhesion. Analytical chemistry

76(18):5257-5264.

[41] Tan JL, Tien J, Pirone DM, Gray DS, Bhadriraju K, & Chen CS (2003) Cells lying on a bed

of microneedles: an approach to isolate mechanical force. Proceedings of the National

Academy of Sciences 100(4):1484-1489.

Page 160: Time-Dependent Mechanics of Living Cells

160

[42] Munevar S, Wang Y, & Dembo M (2001) Traction force microscopy of migrating normal

and H-ras transformed 3T3 fibroblasts. Biophysical journal 80(4):1744-1757.

[43] Wirtz D (2009) Particle-tracking microrheology of living cells: principles and applications.

Annual review of biophysics 38:301-326.

[44] Lim CT, Zhou EH, & Quek ST (2006) Mechanical models for living cells--a review. Journal

of biomechanics 39(2):195-216.

[45] Janmey PA & McCulloch CA (2007) Cell mechanics: integrating cell responses to

mechanical stimuli. Annu. Rev. Biomed. Eng. 9:1-34.

[46] Bao G & Suresh S (2003) Cell and molecular mechanics of biological materials. Nature

materials 2(11):715-725.

[47] Binnig G, Quate CF, & Gerber C (1986) Atomic force microscope. Physical review letters

56(9):930-933.

[48] Morris VJ, Kirby AR, Gunning AP, & World S (1999) Atomic force microscopy for

biologists (Imperial College Press London).

[49] Kunda P, Rodrigues NTL, Moeendarbary E, Liu T, Ivetic A, Charras G, & Baum B (2011)

PP1-Mediated Moesin Dephosphorylation Couples Polar Relaxation to Mitotic Exit.

Current Biology.

[50] Alessandrini A & Facci P (2005) AFM: a versatile tool in biophysics. Measurement science

and technology 16:R65.

[51] Moreno-Flores S, Benitez R, Vivanco MM, & Toca-Herrera JL (2010) Stress relaxation and

creep on living cells with the atomic force microscope: a means to calculate elastic

moduli and viscosities of cell components. Nanotechnology 21:445101.

[52] Moreno-Flores S, Benitez R, Vivanco MM, & Toca-Herrera JL (2010) Stress relaxation

microscopy: Imaging local stress in cells. Journal of biomechanics 43(2):349-354.

[53] Darling EM, Zauscher S, & Guilak F (2006) Viscoelastic properties of zonal articular

chondrocytes measured by atomic force microscopy. Osteoarthritis and Cartilage

14(6):571-579.

[54] Darling EM, Zauscher S, Block JA, & Guilak F (2007) A thin-layer model for viscoelastic,

stress-relaxation testing of cells using atomic force microscopy: do cell properties reflect

metastatic potential? Biophysical journal 92(5):1784-1791.

[55] Kasza KE, Rowat AC, Liu J, Angelini TE, Brangwynne CP, Koenderink GH, & Weitz DA

(2007) The cell as a material. Curr Opin Cell Biol 19(1):101-107.

[56] Fletcher DA & Geissler PL (2009) Active biological materials. Annual review of physical

chemistry 60:469-486.

[57] Julicher F, Kruse K, Prost J, & Joanny JF (2007) Active behavior of the cytoskeleton.

Physics reports 449(1-3):3-28.

[58] Hoffman BD & Crocker JC (2009) Cell mechanics: dissecting the physical responses of cells

to force. Annual review of biomedical engineering 11:259-288.

[59] Mitchison TJ, Charras GT, & Mahadevan L (2008) Implications of a poroelastic cytoplasm

for the dynamics of animal cell shape. Seminars in cell & developmental biology,

(Elsevier), pp 215-223.

[60] Kumar S, Maxwell IZ, Heisterkamp A, Polte TR, Lele TP, Salanga M, Mazur E, & Ingber

DE (2006) Viscoelastic retraction of single living stress fibers and its impact on cell

shape, cytoskeletal organization, and extracellular matrix mechanics. Biophysical journal

90(10):3762-3773.

[61] Bursac P, Lenormand G, Fabry B, Oliver M, Weitz DA, Viasnoff V, Butler JP, & Fredberg

JJ (2005) Cytoskeletal remodelling and slow dynamics in the living cell. Nature materials

4(7):557-561.

[62] Trepat X, Deng L, An SS, Navajas D, Tschumperlin DJ, Gerthoffer WT, Butler JP, &

Fredberg JJ (2007) Universal physical responses to stretch in the living cell. Nature

447(7144):592-595.

[63] Chen DTN, Wen Q, Janmey PA, Crocker JC, & Yodh AG (2010) Rheology of soft

materials. Condensed Matter Physics 1.

[64] Kollmannsberger P & Fabry B (2011) Linear and nonlinear rheology of living cells. Annual

Review of Materials Research 41:75-97.

Page 161: Time-Dependent Mechanics of Living Cells

161

[65] Trepat X, Lenormand G, & Fredberg JJ (2008) Universality in cell mechanics. Soft Matter

4(9):1750-1759.

[66] Ingber DE (2003) Tensegrity I. Cell structure and hierarchical systems biology. Journal of

Cell Science 116(7):1157.

[67] Sollich P (1998) Rheological constitutive equation for a model of soft glassy materials.

Physical Review E 58(1):738.

[68] Bausch AR, Möller W, & Sackmann E (1999) Measurement of local viscoelasticity and

forces in living cells by magnetic tweezers. Biophysical journal 76(1):573-579.

[69] Bausch AR, Hellerer U, Essler M, Aepfelbacher M, & Sackmann E (2001) Rapid stiffening

of integrin receptor-actin linkages in endothelial cells stimulated with thrombin: a

magnetic bead microrheology study. Biophysical journal 80(6):2649-2657.

[70] Dix JA & Verkman AS (2008) Crowding effects on diffusion in solutions and cells. Annu.

Rev. Biophys. 37:247-263.

[71] Mastro AM, Babich MA, Taylor WD, & Keith AD (1984) Diffusion of a small molecule in

the cytoplasm of mammalian cells. Proceedings of the National Academy of Sciences

81(11):3414.

[72] Luby-Phelps K, Castle PE, Taylor DL, & Lanni F (1987) Hindered diffusion of inert tracer

particles in the cytoplasm of mouse 3T3 cells. Proceedings of the National Academy of

Sciences 84(14):4910-4913.

[73] Hoffman BD, Massiera G, Van Citters KM, & Crocker JC (2006) The consensus mechanics

of cultured mammalian cells. Proceedings of the National Academy of Sciences

103(27):10259-10264.

[74] Grünewald K, Medalia O, Gross A, Steven AC, & Baumeister W (2003) Prospects of

electron cryotomography to visualize macromolecular complexes inside cellular

compartments: implications of crowding* 1. Biophysical chemistry 100(1-3):577-591.

[75] Medalia O, Weber I, Frangakis AS, Nicastro D, Gerisch G, & Baumeister W (2002)

Macromolecular architecture in eukaryotic cells visualized by cryoelectron tomography.

Science 298(5596):1209.

[76] Wong IY, Gardel ML, Reichman DR, Weeks ER, Valentine MT, Bausch AR, & Weitz DA

(2004) Anomalous diffusion probes microstructure dynamics of entangled F-actin

networks. Physical review letters 92(17):178101.

[77] Novak IL, Kraikivski P, & Slepchenko BM (2009) Diffusion in cytoplasm: effects of

excluded volume due to internal membranes and cytoskeletal structures. Biophysical

journal 97(3):758-767.

[78] Guigas G, Kalla C, & Weiss M (2007) The degree of macromolecular crowding in the

cytoplasm and nucleoplasm of mammalian cells is conserved. FEBS letters 581(26):5094-

5098.

[79] Alcaraz J, Buscemi L, Grabulosa M, Trepat X, Fabry B, Farré R, & Navajas D (2003)

Microrheology of human lung epithelial cells measured by atomic force microscopy.

Biophysical journal 84(3):2071-2079.

[80] Deng L, Trepat X, Butler JP, Millet E, Morgan KG, Weitz DA, & Fredberg JJ (2006) Fast

and slow dynamics of the cytoskeleton. Nature materials 5(8):636-640.

[81] Gittes F, Schnurr B, Olmsted PD, MacKintosh FC, & Schmidt CF (1997) Microscopic

viscoelasticity: shear moduli of soft materials determined from thermal fluctuations.

Physical review letters 79(17):3286-3289.

[82] Morse DC (1998) Viscoelasticity of concentrated isotropic solutions of semiflexible

polymers. 2. Linear response. Macromolecules 31(20):7044-7067.

[83] Gardel M, Valentine M, Crocker JC, Bausch A, & Weitz DA (2003) Microrheology of

entangled F-actin solutions. Physical review letters 91(15):158302.

[84] Gardel ML, Shin JH, MacKintosh FC, Mahadevan L, Matsudaira P, & Weitz DA (2004)

Elastic behavior of cross-linked and bundled actin networks. Science 304(5675):1301-

1305.

[85] Tharmann R, Claessens M, & Bausch AR (2007) Viscoelasticity of isotropically cross-

linked actin networks. Physical review letters 98(8):88103.

[86] Storm C, Pastore JJ, MacKintosh FC, Lubensky TC, & Janmey PA (2005) Nonlinear

elasticity in biological gels. Nature 435(7039):191-194.

Page 162: Time-Dependent Mechanics of Living Cells

162

[87] Rubenstein M & Colby RH (2003) Polymer Physics. (Oxford Univ. Press: New York).

[88] Wilhelm J & Frey E (2003) Elasticity of stiff polymer networks. Physical review letters

91(10):108103.

[89] Head DA, Levine AJ, & MacKintosh FC (2003) Deformation of cross-linked semiflexible

polymer networks. Physical review letters 91(10):108102.

[90] Levine AJ & Lubensky T (2001) Response function of a sphere in a viscoelastic two-fluid

medium. Physical Review E 63(4):041510.

[91] De Gennes PG, Pincus P, Velasco R, & Brochard F (1976) Remarks on polyelectrolyte

conformation. Journal de physique 37(12):1461-1473.

[92] Xu K, Babcock HP, & Zhuang X (2012) Dual-objective STORM reveals three-dimensional

filament organization in the actin cytoskeleton. Nature Methods.

[93] Gardel M, Nakamura F, Hartwig J, Crocker J, Stossel T, & Weitz D (2006) Prestressed F-

actin networks cross-linked by hinged filamins replicate mechanical properties of cells.

Proceedings of the National Academy of Sciences of the United States of America

103(6):1762-1767.

[94] Gardel ML, Nakamura F, Hartwig J, Crocker JC, Stossel TP, & Weitz DA (2006) Stress-

dependent elasticity of composite actin networks as a model for cell behavior. Physical

review letters 96(8):88102.

[95] Ingber DE (1993) Cellular tensegrity: defining new rules of biological design that govern the

cytoskeleton. Journal of Cell Science 104:613-613.

[96] Joanny JF & Prost J (2009) Active gels as a description of the actin‐myosin cytoskeleton.

HFSP journal 3(2):94-104.

[97] Liverpool T & Marchetti MC (2007) Bridging the microscopic and the hydrodynamic in

active filament solutions. EPL (Europhysics Letters) 69(5):846.

[98] Liverpool T & Marchetti MC (2006) Rheology of active filament solutions. Physical review

letters 97(26):268101.

[99] Ball P (2008) Water as an active constituent in cell biology. Chemical reviews 108(1):74-

108.

[100] Charras GT, Yarrow JC, Horton MA, Mahadevan L, & Mitchison TJ (2005) Non-

equilibration of hydrostatic pressure in blebbing cells. Nature 435(7040):365-369.

[101] Charras GT, Mitchison TJ, & Mahadevan L (2009) Animal cell hydraulics. Journal of Cell

Science 122(18):3233-3241.

[102] Charras GT, Coughlin M, Mitchison TJ, & Mahadevan L (2008) Life and times of a

cellular bleb. Biophysical journal 94(5):1836-1853.

[103] Rosenbluth MJ, Crow A, Shaevitz JW, & Fletcher DA (2008) Slow stress propagation in

adherent cells. Biophysical journal 95(12):6052-6059.

[104] Zhang W & Robinson DN (2005) Balance of actively generated contractile and resistive

forces controls cytokinesis dynamics. Proceedings of the National Academy of Sciences

of the United States of America 102(20):7186.

[105] Dembo M & Harlow F (1986) Cell motion, contractile networks, and the physics of

interpenetrating reactive flow. Biophysical journal 50(1):109-121.

[106] Wang H (2000) Theory of linear poroelasticity: with applications to geomechanics and

hydrogeology (Princeton Univ Pr).

[107] Charras G & Paluch E (2008) Blebs lead the way: how to migrate without lamellipodia.

Nature Reviews Molecular Cell Biology 9(9):730-736.

[108] Lai WM, Rubin D, & Krempl E (2009) Introduction to continuum mechanics (Butterworth-

Heinemann).

[109] Landau L & Lifshitz E (1959) Theory of elasticity. Pergamon Press Ltd., Oxford, U.K 30–

36.

[110] Johnson KL (1982) One hundred years of Hertz contact. ARCHIVE: Proceedings of the

Institution of Mechanical Engineers 1847-1982 (vols 1-196) 196(1982):363-378.

[111] Christensen RM & Freund LB (1971) Theory of viscoelasticity. Journal of Applied

Mechanics 38:720.

[112] Fabry B, Maksym GN, Butler JP, Glogauer M, Navajas D, Taback NA, Millet EJ, &

Fredberg JJ (2003) Time scale and other invariants of integrative mechanical behavior in

living cells. Physical Review E 68(4):041914.

Page 163: Time-Dependent Mechanics of Living Cells

163

[113] Heymans N (2004) Fractional calculus description of non-linear viscoelastic behaviour of

polymers. Nonlinear dynamics 38(1):221-231.

[114] Mainardi F (2012) Fractional calculus: some basic problems in continuum and statistical

mechanics. Arxiv preprint arXiv:1201.0863.

[115] Detournay E & Cheng AHD (1993) Fundamentals of poroelasticity. (Pergamon).

[116] Biot MA (1941) General theory of three-dimensional consolidation. Journal of applied

physics 12(2):155-164.

[117] Mow VC, Kuei SC, Lai WM, & Armstrong CG (1980) Biphasic creep and stress relaxation

of articular cartilage in compression: theory and experiments. Journal of biomechanical

engineering 102:73.

[118] Scheidegger AE (1958) The physics of flow through porous media. Soil Science 86(6):355.

[119] Yoon J, Cai S, Suo Z, & Hayward RC (2010) Poroelastic swelling kinetics of thin hydrogel

layers: comparison of theory and experiment. Soft Matter.

[120] Hu Y, Zhao X, Vlassak JJ, & Suo Z (2010) Using indentation to characterize the

poroelasticity of gels. Applied Physics Letters 96:121904.

[121] Chan EP, Hu Y, Johnson PM, Suo Z, & Stafford CM (2012) Spherical indentation testing

of poroelastic relaxations in thin hydrogel layers. Soft Matter 8(5):1492-1498.

[122] Ory DS, Neugeboren BA, & Mulligan RC (1996) A stable human-derived packaging cell

line for production of high titer retrovirus/vesicular stomatitis virus G pseudotypes.

Proceedings of the National Academy of Sciences of the United States of America

93(21):11400.

[123] Riedl J, Crevenna AH, Kessenbrock K, Yu JH, Neukirchen D, Bista M, Bradke F, Jenne D,

Holak TA, & Werb Z (2008) Lifeact: a versatile marker to visualize F-actin. Nature

Methods 5(7):605-607.

[124] Bancaud A, Huet S, Daigle N, Mozziconacci J, Beaudouin J, & Ellenberg J (2009)

Molecular crowding affects diffusion and binding of nuclear proteins in heterochromatin

and reveals the fractal organization of chromatin. The EMBO journal 28(24):3785-3798.

[125] Zhou EH, Trepat X, Park CY, Lenormand G, Oliver MN, Mijailovich SM, Hardin C, Weitz

DA, Butler JP, & Fredberg JJ (2009) Universal behavior of the osmotically compressed

cell and its analogy to the colloidal glass transition. Proceedings of the National Academy

of Sciences 106(26):10632-10637.

[126] Hoffmann EK, Lambert IH, & Pedersen SF (2009) Physiology of cell volume regulation in

vertebrates. Physiological reviews 89(1):193-277.

[127] Demaison C, Parsley K, Brouns G, Scherr M, Battmer K, Kinnon C, Grez M, & Thrasher

AJ (2002) High-level transduction and gene expression in hematopoietic repopulating

cells using a human imunodeficiency virus type 1-based lentiviral vector containing an

internal spleen focus forming virus promoter. Human gene therapy 13(7):803-813.

[128] Zheng Y, Wong ML, Alberts B, & Mitchison T (1995) Nucleation of microtubule assembly

by a gamma-tubulin-containing ring complex. Nature 378:578-583.

[129] Keppler A & Ellenberg J (2009) Chromophore-assisted laser inactivation of -and -tubulin

SNAP-tag fusion proteins inside living cells. ACS Chemical Biology 4(2):127-138.

[130] Werner NS, Windoffer R, Strnad P, Grund C, Leube RE, & Magin TM (2004)

Epidermolysis bullosa simplex-type mutations alter the dynamics of the keratin

cytoskeleton and reveal a contribution of actin to the transport of keratin subunits.

Molecular biology of the cell 15(3):990-1002.

[131] Long R, Hall MS, Wu M, & Hui CY (2011) Effects of gel thickness on microscopic

indentation measurements of gel modulus. Biophysical journal 101(3):643-650.

[132] Gavara N & Chadwick RS (2012) Determination of the elastic moduli of thin samples and

adherent cells using conical atomic force microscope tips. Nature nanotechnology

7(11):733-736.

[133] Lin DC, Dimitriadis EK, & Horkay F (2007) Robust strategies for automated AFM force

curve analysis—I. Non-adhesive indentation of soft, inhomogeneous materials. Journal of

biomechanical engineering 129:430.

[134] Charras GT, Hu CK, Coughlin M, & Mitchison TJ (2006) Reassembly of contractile actin

cortex in cell blebs. The Journal of cell biology 175(3):477-490.

Page 164: Time-Dependent Mechanics of Living Cells

164

[135] Hui CY, Lin YY, Chuang FC, Shull KR, & Lin WC (2006) A contact mechanics method

for characterizing the elastic properties and permeability of gels. Journal of Polymer

Science Part B: Polymer Physics 44(2):359-370.

[136] Hu Y, Chen X, Whitesides GM, Vlassak JJ, & Suo Z (2011) Indentation of

polydimethylsiloxane submerged in organic solvents. Journal of Materials Research

26(06):785-795.

[137] Kalcioglu ZI, Mahmoodian R, Hu Y, Suo Z, & Van Vliet KJ (2012) From macro-to

microscale poroelastic characterization of polymeric hydrogels via indentation. Soft

Matter 8:3393-3398.

[138] Speidel M, Jonáš A, & Florin EL (2003) Three-dimensional tracking of fluorescent

nanoparticles with subnanometer precision by use of off-focus imaging. Optics letters

28(2):69-71.

[139] Axelrod D, Koppel D, Schlessinger J, Elson E, & Webb W (1976) Mobility measurement

by analysis of fluorescence photobleaching recovery kinetics. Biophysical journal

16(9):1055-1069.

[140] Soumpasis D (1983) Theoretical analysis of fluorescence photobleaching recovery

experiments. Biophysical journal 41(1):95-97.

[141] Kang M, Day CA, Drake K, Kenworthy AK, & DiBenedetto E (2009) A generalization of

theory for two-dimensional fluorescence recovery after photobleaching applicable to

confocal laser scanning microscopes. Biophysical journal 97(5):1501-1511.

[142] Chan EP, Deeyaa B, Johnson PM, & Stafford CM (2012) Poroelastic relaxation of

polymer-loaded hydrogels. Soft Matter 8(31):8234-8240.

[143] Fredberg JJ & Stamenovic D (1989) On the imperfect elasticity of lung tissue. Journal of

applied physiology 67(6):2408-2419.

[144] Smith BA, Tolloczko B, Martin JG, & Grütter P (2005) Probing the viscoelastic behavior

of cultured airway smooth muscle cells with atomic force microscopy: stiffening induced

by contractile agonist. Biophysical journal 88(4):2994-3007.

[145] Okajima T, Tanaka M, Tsukiyama S, Kadowaki T, Yamamoto S, Shimomura M, &

Tokumoto H (2007) Stress relaxation of HepG2 cells measured by atomic force

microscopy. Nanotechnology 18:084010.

[146] Wottawah F, Schinkinger S, Lincoln B, Ebert S, Müller K, Sauer F, Travis K, & Guck J

(2005) Characterizing single suspended cells by optorheology. Acta biomaterialia

1(3):263-271.

[147] Desprat N, Richert A, Simeon J, & Asnacios A (2005) Creep function of a single living

cell. Biophysical journal 88(3):2224-2233.

[148] Wottawah F, Schinkinger S, Lincoln B, Ananthakrishnan R, Romeyke M, Guck J, & Käs J

(2005) Optical rheology of biological cells. Physical review letters 94(9):98103.

[149] Lenormand G, Millet E, Fabry B, Butler JP, & Fredberg JJ (2004) Linearity and time-scale

invariance of the creep function in living cells. Journal of The Royal Society Interface

1(1):91-97.

[150] Keren K, Yam PT, Kinkhabwala A, Mogilner A, & Theriot JA (2009) Intracellular fluid

flow in rapidly moving cells. Nature cell biology 11(10):1219-1224.

[151] Moreno-Flores S, Benitez R, & Toca-Herrera JL (2010) Stress relaxation and creep on

living cells with the atomic force microscope: a means to calculate elastic moduli and

viscosities of cell components. Nanotechnology 21:445101.

[152] Skotheim J & Mahadevan L (2004) Dynamics of poroelastic filaments. Proceedings of the

Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences

460(2047):1995-2020.

[153] Avril S, Schneider F, Boissier C, & Li ZY (2011) In vivo velocity vector imaging and time-

resolved strain rate measurements in the wall of blood vessels using MRI. Journal of

biomechanics 44(5):979-983.

[154] Li P, Liu A, Shi L, Yin X, Rugonyi S, & Wang RK (2011) Assessment of strain and strain

rate in embryonic chick heart in vivo using tissue Doppler optical coherence tomography.

Physics in medicine and biology 56:7081–7092.

[155] Perlman CE & Bhattacharya J (2007) Alveolar expansion imaged by optical sectioning

microscopy. Journal of applied physiology 103(3):1037-1044.

Page 165: Time-Dependent Mechanics of Living Cells

165

[156] Ibata K, Takimoto S, Morisaku T, Miyawaki A, & Yasui M (2011) Analysis of Aquaporin-

Mediated Diffusional Water Permeability by Coherent Anti-Stokes Raman Scattering

Microscopy. Biophysical journal 101(9):2277-2283.

[157] Fels J, Orlov SN, & Grygorczyk R (2009) The hydrogel nature of mammalian cytoplasm

contributes to osmosensing and extracellular pH sensing. Biophysical journal

96(10):4276-4285.

[158] Sehy JV, Banks AA, Ackerman JJH, & Neil JJ (2002) Importance of intracellular water

apparent diffusion to the measurement of membrane permeability. Biophysical journal

83(5):2856-2863.

[159] Potma EO, De Boeij WP, van Haastert PJM, & Wiersma DA (2001) Real-time

visualization of intracellular hydrodynamics in single living cells. Proceedings of the

National Academy of Sciences 98(4):1577.

[160] Lang F, Busch GL, Ritter M, Volkl H, Waldegger S, Gulbins E, & Haussinger D (1998)

Functional significance of cell volume regulatory mechanisms. Physiological reviews

78(1):247.

[161] Derfus AM, Chan WCW, & Bhatia SN (2004) Intracellular delivery of quantum dots for

live cell labeling and organelle tracking. Advanced Materials 16(12):961-966.

[162] Swaminathan R, Bicknese S, Periasamy N, & Verkman AS (1996) Cytoplasmic viscosity

near the cell plasma membrane: translational diffusion of a small fluorescent solute

measured by total internal reflection-fluorescence photobleaching recovery. Biophysical

Journal 71(2):1140-1151.

[163] Swaminathan R, Hoang CP, & Verkman A (1997) Photobleaching recovery and anisotropy

decay of green fluorescent protein GFP-S65T in solution and cells: cytoplasmic viscosity

probed by green fluorescent protein translational and rotational diffusion. Biophysical

journal 72(4):1900-1907.

[164] Kao HP, Abney JR, & Verkman AS (1993) Determinants of the translational mobility of a

small solute in cell cytoplasm. The Journal of cell biology 120(1):175-184.

[165] Phair RD & Misteli T (2000) High mobility of proteins in the mammalian cell nucleus.

Nature 404(6778):604-608.

[166] Phillips RJ (2000) A hydrodynamic model for hindered diffusion of proteins and micelles

in hydrogels. Biophysical Journal 79(6):3350.

[167] Happel J & Brenner H (1983) Low Reynolds number hydrodynamics: with special

applications to particulate media (Springer).

[168] Persson E & Halle B (2008) Cell water dynamics on multiple time scales. Proceedings of

the National Academy of Sciences 105(17):6266.

[169] Rotsch C & Radmacher M (2000) Drug-induced changes of cytoskeletal structure and

mechanics in fibroblasts: an atomic force microscopy study. Biophysical journal

78(1):520-535.

[170] Rotsch C, Jacobson K, & Radmacher M (1999) Dimensional and mechanical dynamics of

active and stable edges in motile fibroblasts investigated by using atomic force

microscopy. Proceedings of the National Academy of Sciences of the United States of

America 96(3):921.

[171] Wakatsuki T, Schwab B, Thompson NC, & Elson EL (2001) Effects of cytochalasin D and

latrunculin B on mechanical properties of cells. Journal of cell science 114(Pt 5):1025.

[172] Low SH, Mukhina S, Srinivas V, Ng CZ, & Murata-Hori M (2010) Domain analysis of α-

actinin reveals new aspects of its association with F-actin during cytokinesis.

Experimental cell research 316(12):1925-1934.

[173] Shu HB & Joshi HC (1995) Gamma-tubulin can both nucleate microtubule assembly and

self-assemble into novel tubular structures in mammalian cells. The Journal of cell

biology:1137-1147.

[174] Spagnoli C, Beyder A, Besch S, & Sachs F (2008) Atomic force microscopy analysis of

cell volume regulation. Physical Review E 78(3):31916.

[175] Albrecht-Buehler G & Bushnell A (1982) Reversible compression of cytoplasm.

Experimental cell research 140(1):173-189.

[176] Mizuno D, Tardin C, Schmidt CF, & MacKintosh FC (2007) Nonequilibrium mechanics of

active cytoskeletal networks. Science 315(5810):370-373.

Page 166: Time-Dependent Mechanics of Living Cells

166

[177] Stewart MP, Helenius J, Toyoda Y, Ramanathan SP, Muller DJ, & Hyman AA (2011)

Hydrostatic pressure and the actomyosin cortex drive mitotic cell rounding. Nature

469(7329):226-230.

[178] Stamenovic D, Rosenblatt N, Montoya-Zavala M, Matthews BD, Hu S, Suki B, Wang N, &

Ingber DE (2007) Rheological behavior of living cells is timescale-dependent.

Biophysical journal 93(8):L39-L41.

[179] Van Citters KM, Hoffman BD, Massiera G, & Crocker JC (2006) The role of F-actin and

myosin in epithelial cell rheology. Biophysical journal 91(10):3946-3956.

[180] Mukhina S, Wang Y, & Murata-Hori M (2007) [alpha]-Actinin Is Required for Tightly

Regulated Remodeling of the Actin Cortical Network during Cytokinesis. Developmental

cell 13(4):554-565.

[181] Ando T & Skolnick J (2010) Crowding and hydrodynamic interactions likely dominate in

vivo macromolecular motion. Proceedings of the National Academy of Sciences

107(43):18457-18462.

[182] Gittes F, Mickey B, Nettleton J, & Howard J (1993) Flexural rigidity of microtubules and

actin filaments measured from thermal fluctuations in shape. The Journal of cell biology

120(4):923-934.

[183] Zicha D, Dobbie IM, Holt MR, Monypenny J, Soong DYH, Gray C, & Dunn GA (2003)

Rapid actin transport during cell protrusion. Science 300(5616):142-145.

[184] Pollard TD & Borisy GG (2003) Cellular motility driven by assembly and disassembly of

actin filaments. Cell 112(4):453-465.

[185] De Gennes PG (1976) Dynamics of entangled polymer solutions ( I&II). Macromolecules

9(4):587-598.

[186] Hawkins RJ, Piel M, Faure-Andre G, Lennon-Dumenil A, Joanny J, Prost J, & Voituriez R

(2009) Pushing off the walls: a mechanism of cell motility in confinement. Physical

review letters 102(5):58103.